You are on page 1of 17

Journal of Earthquake and Tsunami

(2019)
.c World Scienti¯c Publishing Company
#
DOI: 10.1142/S1793431119500040

Critical Resistance A®ecting Sub- to Super-Critical


Transition Flow by Vegetation
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

Ghufran Ahmed Pasha*,†,§ and Norio Tanaka*,‡,¶


*Graduate School of Science and Engineering, Saitama University
255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan
J. Earthquake and Tsunami Downloaded from www.worldscientific.com


Department of Civil Engineering
University of Engineering and Technology, Taxila 47050, Pakistan

International Institute for Resilient Society, Saitama University
255 Shimo-okubo, Sakura-ku, Saitama 338-8570, Japan
§ghufran.ahmed@uettaxila.edu.pk

tanaka01@mail.saitama-u.ac.jp

Received 2 March 2018


Accepted 9 January 2019
Published 7 March 2019

In order to design a vegetation structure to mitigate °oods resulting from extreme events like
tsunamis, vegetation density and thickness (width) are important parameters. Flow passing
through vegetation faces great resistance, which results in a backwater rise on upstream (U/S)
vegetation, increases the water slope inside the vegetation, and for some cases, forms a hydraulic
jump downstream (D/S) of the vegetation, thus transforming a subcritical °ow to supercritical
[Pasha, G. A. and Tanaka, N. [2017] \Undular hydraulic jump formation and energy loss in a
°ow through emergent vegetation of varying thickness and density," Ocean Eng. 141, 308–
325.]. Like the concepts of critical velocity and critical slope, this paper introduces the concept of
\critical resistance of vegetation," which is de¯ned as \resistance o®ered by vegetation that
transforms a subcritical °ow to supercritical." An analytical approach to ¯nd the water depths
U/S, inside, and D/S of vegetation is introduced and validated well by laboratory experiments.
Critical resistance was determined against vegetation of variable densities (G=d, where G ¼
spacing of each cylinder in the cross-stream direction, d ¼ diameter of the cylinder), thicknesses
(dn, where d ¼ diameter of a cylinder and n ¼ number of cylinders in a stream-wise direction
per unit of cross-stream width), and the initial Froude number (Fro ). A subcritical °ow
(Fro ¼ 0:55  0:75, without vegetation) was transformed to a supercritical °ow (D/S vege-
tation) with a range of Froude numbers of 1.6–1.9, 1.1–1.2, and 0.85–0.98 against G=d ratios of
0.25, 1.09, and 2.13, respectively, thus de¯ning G=d ¼ 1:0 as the critical resistance. However,
altering vegetation thickness did not change the results.

Keywords: Vegetation; critical resistance; subcritical °ow; supercritical °ow; backwater rise;
undular hydraulic jump.

¶ Corresponding author.

1950004-1
G. A. Pasha & N. Tanaka

1. Introduction
The e®ectiveness of coastal vegetation in mitigating a tsunami has been recognized
by many studies since the 1998 Papua New Guinea tsunami [Dengler and Preuss, 2003;
Hiraishi and Harada, 2003], 2004 Indian Ocean tsunami [Danielsen et al., 2005;
Mascarenhas and Jayakumar, 2008; Tanaka et al., 2007], and 2011 Great East Japan
tsunami [Nandasena et al., 2012; Tanaka et al., 2014]. As shown by previous research
[Shuto, 1987; Tanaka et al., 2007], coastal vegetation plays a vital role in tsunami
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

mitigation by trapping debris, providing a soft-landing place or escape route, and


dissipating energy. Arti¯cial methods like construction of sea walls and embank-
ments, installment of tsunami gates, and erection of breakwater structures could
prove to be di±cult for developing countries to implement because they require
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

enormous capital investment [Tanaka, 2009] and have greater chances of failure if a
large tsunami that exceeds the capacity of the arti¯cial structure occurs. On other
hand, natural methods such as coastal forests are considered to be e®ective barriers
to mitigate tsunami damage from both economic and environmental points of view
[Kathiresan and Rajendran, 2005; Osti et al., 2009; Tanaka, 2009; Yanagisawa et al.,
2009, 2010].
Attempts to evaluate the e®ectiveness of a coastal forest have been conducted by
laboratory experiments [Harada and Imamura, 2000; Iimura and Tanaka, 2012;
Pasha and Tanaka, 2016a, 2016b, 2017; Pasha et al., 2018] and also by numerical
simulations by changing the tsunami and forest characteristics [Hiraishi and Harada,
2003; Harada and Imamura, 2005; Nandasena et al., 2008; Iimura and Tanaka, 2013].
The e®ectiveness of coastal forests against tsunamis was also investigated by
Kathiresan and Rajendran [2005]; Irtem et al. [2009]; Yanagisawa et al. [2009], and
Nateghi et al. [2016]. On the other hand, the limitations of the tsunami mitigation
capacity of a coastal forest were also discussed in terms of destruction of the coastal
forest itself [Tanaka et al., 2007; Huang et al., 2013], production of driftwood
[Dengler and Preuss, 2003; Cochard et al., 2008], de¯ciencies in coastal forests such as
open gaps [Mascarenhas and Jayakumar, 2008; Thuy et al., 2009; Nandasena et al.,
2012], the breaking moment of the trees [Tanaka et al., 2013], and the magnitude of
the tsunami [Tanaka, 2009].
During the extensive damage of the 2011 Great East Japan tsunami (GEJT),
coastal forests played a vital role in tsunami mitigation. If the Froude number is
relatively lower on the inland side, then the trees can withstand the pressure of
°oating debris and can trap debris [Tanaka, 2012; Pasha and Tanaka, 2016a; Tanaka
and Onai, 2017; Tanaka and Ogino, 2017]. This was also con¯rmed at Sendai, Natori,
Yotsukura Beach, and Shinmaiko Beach by post tsunami surveys of GEJT. The
Reconstruction Agency [2011] has recommended taking advantage of coastal forests
when reconstructing coastal areas in the future. The force of the water passing
through the vegetation decreases, minimizing the damage behind the vegetation.
Iimura and Tanaka [2012] investigated the e®ects of vegetation density both
experimentally and analytically and con¯rmed that both the level and velocity of

1950004-2
Critical Resistance of Vegetation

water behind the vegetation are reduced considerably by increasing the density of
vegetation. However, as seen from the previous studies [Pasha and Tanaka, 2017;
Pasha et al., 2018], an undular hydraulic jump forms at the downstream vegetation.
An undular hydraulic jump is one type of jump that forms under supercritical
conditions but a low Froude number (Fr ¼ u/(gh)0:5 , u: velocity, g: gravitational
acceleration, h: water depth) where Fr is slightly larger than one. However, by
increasing the Froude number and making the vegetation denser or by changing any
other condition, the intensity of a hydraulic jump can be increased, which can lead to
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

local scouring despite the low energy dissipation by the jump itself. Excessive
scouring can also undermine the roots of vegetation, which may result in possible
failure of protection by vegetation. Moreover, if the hydraulic jump is long, the
supercritical °ow region extends farther, which can be harmful to inland residential
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

areas. In order to prevent this, knowledge of the °ow structure behind the vegetation
must be investigated so that possible remedial measures can be proposed.
Depending upon the vegetation conditions (density and thickness), the °ow can
be transformed to a supercritical °ow as it passes through the vegetation. In this
study, the critical resistance of vegetation is de¯ned as \resistance o®ered by vege-
tation, for which the subcritical °ow (without vegetation and upstream of
vegetation) becomes supercritical (downstream of vegetation)" [Fig. 1(a)]. For

(a)

(b) (c)

Fig. 1. (a) Water pro¯le of vegetation, and de¯nition of critical resistance of vegetation, where HGL is
the hydraulic grade line, (b) plan view, (c) cross-sectional view.

1950004-3
G. A. Pasha & N. Tanaka

reconstructing coastal areas in the future, the proper layout of vegetation and
optimum spacing of trees are important parameters for consideration. The °ow
resistance varies with the °ow velocity, concentration of trees, their diameters, and
elastic properties of the individual trees [Noarayanan et al., 2012]. Flexible emergent
vegetation o®ers great resistance to the °ow and quickly reduces the average velocity
[Li et al., 2015], and the °ow resistance greatly depends on the volume of vegetation
[Yokojima et al., 2015]. The tree spacing should not be so large that the °ow can
easily pass through it, nor so dense that it creates a supercritical °ow covering a wide
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

portion of ground in the downstream vegetation. Pasha and Tanaka [2017] revealed
that the denser the vegetation, the higher the energy dissipation. The concept of
designing dense vegetation that produces a hydraulic jump downstream for miti-
gation can be utilized when available space is limited. Therefore, it is important to
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

know whether the downstream °ow is subcritical or supercritical so that the distance
at which inland residential construction is safe can be determined.
Therefore, the objective of the current study was to characterize the changes in
water pro¯les (upstream, inside, and downstream of vegetation) with reference to
changing vegetation and °ow conditions. In addition, to classify the vegetation
conditions that transform the °ow from sub- to supercritical, an analytical method is
introduced to determine water depths. The results of the analytical method are
validated by the experimental results in Pasha and Tanaka [2017].

2. Analytical Method to Find Water Depths


Momentum analysis is a useful method for solving °ow problems involving energy
loss due to an obstruction such as a bridge pier. The obstruction o®ers resistance to
°ow causing a backwater rise upstream. According to Sturm [2001], the two types of
°ow with an obstruction are Type I and Type II. Type I is a subcritical approach °ow
with a decrease in water depth when passing through a constriction. In Type II °ow,
choking occurs when the critical depth exists in the constriction. To ¯nd the water
depth in the upstream vegetation analytically, the momentum analysis approach of
Type I °ow can be used because the °ow is subcritical without vegetation.
Figure 1(a) shows the water pro¯le through vegetation. Without vegetation, the
initial water depth is ho , and is assumed to rise to h1 when vegetation is added to the
bed. Pasha and Tanaka [2017] measured the backwater rise due to vegetation (G=d
0.25, 1.09, 2.13 and dn 180, 380, 580 No.cm) experimentally and found that the
backwater rise increases with increases in vegetation density and thickness. (The
details of G, d, and dn are provided in Sec. 2.2). The momentum function is de¯ned as:

Q2 h q2 b2
M ¼ Ahc þ ¼ ðbhÞ þ ð1Þ
gA 2 gðbhÞ
where hc is the distance below the free surface to the centroid of the area, A is the area
(for a rectangular channel, area (AÞ ¼ channel width (bÞ x water depth (hÞÞ, Q is the
discharge (m3/s), and q is the unit discharge (m2/s). For a unit strip, i.e., b ¼ 1, and

1950004-4
Critical Resistance of Vegetation

thus the momentum function becomes:


h2 q2
M¼ þ ð2Þ
2 gh
The momentum equation can be written between Sections 1 and 3 of Fig. 1(a), to give:
M1 ¼ M3 þ FD = ð3Þ
where M1 and M3 are the momentum function values at Secs. 1 and 3, respectively, FD
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

is the drag force exerted by trees in the vegetation model, and  is the speci¯c weight of
water. It is assumed that the backwater rise is due only to the resistance to °ow o®ered
by trees because the e®ect of side walls and channel slope (1/500) is extremely small
and therefore neglected in the current analytical approach to ¯nd backwater rise. In the
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

study conducted by Pasha and Tanaka [2017], an undular hydraulic jump downstream
of the vegetation was observed that contributed to energy loss to some extent. How-
ever, after the undular hydraulic jump, the water depth became almost equal to the
initial water depth (ho Þ without vegetation. Thus, the water depth y3 in Fig. 1(a) is
assumed to be almost equal to the initial water depth (ho Þ and the Froude number (Fr3 Þ
equal to Fro . The change in depth or backwater rise h ¼ ðh1  y3 or h1  ho Þ can be
determined against the known conditions at downstream Section 3 (since y3  ho Þ and
with the drag force given as:
FD ¼ CD At V 12 =2; ð4Þ
where At is the frontal area of the tree, CD is the drag coe±cient (depends upon the °ow
and vegetation conditions), and V1 is the velocity.

2.1. Frontal area of trees


The frontal area (At Þ is the summation of areas of all the trees in a unit strip.
From Fig. 1(b), the frontal area is de¯ned as:
 
d
At ¼ m h ð5Þ
S 1

Here, m is the number of rows of trees in the vegetation model, d is the diameter of a
tree, and S is the spacing between trees (center to center) [Fig. 1(b)]. To ¯nd the
water depth at the front of all rows of cylinders except the last row, tree
spacing becomes half, i.e. S ¼ D/2 due to the staggered arrangement of trees, where
D is the distance between cylinders [Fig. 1(b)]. Because the frontal area of the trees is
a summation of all the trees in a unit strip, it is multiplied by the number of rows
of trees (mÞ. In the reference ¯gure [Figs. 1(b) and 1(c), if, for example, the number
of rows is 5, then m ¼ 5. However, to ¯nd the water depth at the front of the last
row of a vegetation model, tree spacing is S ¼ D and m ¼ 1. The number of rows (m)
can be calculated with reference to vegetation width (W ) and cylinders spacing

1950004-5
G. A. Pasha & N. Tanaka

ðDÞ as:

2W
m ¼ pffiffiffi : ð6Þ
3D
Substituting values in Eq. (3) for a rectangular channel, we have:
h 21 q2 y2 q2 mCD dh1 V 12
þ ¼ 3þ þ : ð7Þ
2 gh1 2 gy3 2gS
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

Referring to the study of Sturm [2001] Eq. (7) can be non-dimensionalized in terms of
the downstream Froude number, Fr3 , to produce:
 
3 þ 32 þ 2  2Fr 23   ’Fr 23 ¼ 0 ð8Þ
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

In Eq. (8),  ¼ h=y3 , which is the ratio of backwater rise to downstream water
depth where y3 and Fr3 are equal to ho and Fro , respectively, as explained above.
Coe±cient ’ is de¯ned as:
’ ¼ m  CD d=S: ð9Þ

Using Newton's method, Eq. (8) was solved to get the backwater rise against dif-
ferent vegetation and °ow conditions. In order to accurately ¯nd the water depths,
the drag coe±cient (CD Þ must be known precisely because it greatly a®ects the
accuracy of the analytical method. Pasha and Tanaka [2017] measured the drag
coe±cient of an individual cylinder at the front and back of vegetation. In the current
study, the mean values of CD ¼ 2:0, 1.3, and 1.0, which were found experimentally
by Pasha and Tanaka [2017], were used against dense, intermediate, and sparse
vegetation arrangements, respectively.

2.2. Vegetation and °ow conditions


In order to validate the analytical model, the vegetation and °ow conditions were
kept the same as the conditions that prevailed in the experimental investigation by
Pasha and Tanaka [2017]. This study de¯ned the initial Froude number (Fro Þ when
the reference velocity and water depth were used without a vegetation model in a
water °ume. Steady subcritical conditions were set in the experiments of Pasha and
Tanaka [2017]. For creating subcritical conditions of an inundating tsunami, the
water depths without a vegetation model [ho in Fig. 1(a)] in the experiment were
from 3 to 7 cm in increments of 0.5 cm, setting the initial Froude number approxi-
mately equal to 0.55–0.75. The tree species selected for the vegetation model was the
Japanese pine tree (average tree height ¼ 15 m and trunk diameter ¼ 0:4m) found
on the Sendai Plain; its tree crown was high relative to the tsunami height, and the
trees can be considered as circular cylinders [Tanaka et al., 2014]. For a 1/100 scale
model, wooden cylinders with a diameter of 4 mm were used as tree models in a
staggered arrangement. Many studies have clari¯ed the °ow behavior of staggered

1950004-6
Critical Resistance of Vegetation

and grid arrangements. The drag force behaves di®erently in these two arrange-
ments. According to Takemura and Tanaka [2007], the drag force o®ered by a colony
model is greater in a staggered arrangement as compared to a grid arrangement.
Bokaian and Geoola [1984] discussed the variation of the drag coe±cient as a function
of longitudinal and lateral spacing between two cylinders. In Fig. 1(b), if \D" is the
spacing of each cylinder and `d' the diameter of cylinders, then according to Bokaian
and Geoola [1984] the drag coe±cient of a circular cylinder (Cd Þ at the back was smaller
than that at the front and recovered about 98% of the Cd on the front cylinder in the
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

distance of D=d > 20 when the two cylinders were arranged in a streamwise direction
(grid arrangement). In the staggered arrangement, on the other hand, the Cd of the
backside cylinder took about 98% value even at the close distance of D=d ¼ 1. Thus, we
assumed that the staggered arrangement is more e®ective than the grid arrangement
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

and was therefore considered in this study. Figure 1(b) and Table 1 show details of the
vegetation arrangement, where D is the distance between cylinders and W is the width
of the vegetation model.
In order to investigate the energy loss through vegetation, Pasha and
Tanaka [2017] selected vegetation models with three di®erent G=d values (0.25, 1.09,
and 2.13) where G represents the clear spacing between each cylinder in a cross-
stream direction, and d is the diameter of a cylinder [Fig. 1(b)]. A model with G=d of
0.25 represents dense vegetation, while G=d of 1.09 and 2.13 characterize
intermediate and sparse vegetations, respectively. According to Takemura and
Tanaka [2007], °ow structures di®er depending on the G=d arrangement of the
vegetation model (vegetation density). In the ¯eld, sparse vegetation (G=d 2.13) can
be planted using Japanese red pine (Pinus densi°ora) and black pine (Pinus thun-
bergii) trees. However, small G=d values like intermediate (G=d 1.09) and dense (G=d
0.25) vegetation do not correspond to real forests in the area of applicability. D and
W were determined under the same vegetation thickness, dn (No.cm), which is
expressed as a function of summed tree diameters and is de¯ned as a product of the
diameter of a tree ðdÞ at breast height and the number of trees (n) in a rectangle with

Table 1. Flow and vegetation conditions.

Vegetation Vegetation
Case density D W thickness Vegetation
no. Initial froude number \Fr" o \G/d " (cm) (cm) \dn" (No. cm) type

1 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 0.25 1 3.88 179.21 Dense
2 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 1.09 1.67 10.55 174.72 Intermediate
3 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 2.13 2.5 24.27 179.36 Sparse
4 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 0.25 1 8.23 380.13 Dense
5 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 1.09 1.67 23.60 390.85 Intermediate
6 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 2.13 2.5 52.48 387.83 Sparse
7 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 0.25 1 12.58 581.05 Dense
8 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 1.09 1.67 35.2 582.96 Intermediate
9 0.57, 0.62, 0.65, 0.66, 0.68, 0.69, 0.70, 0.71, 0.73 2.13 2.5 78.52 580.27 Sparse

1950004-7
G. A. Pasha & N. Tanaka

a frontage of unit length along the shoreline and depth equal to the width of the
forest (W ) [Fig. 1(b)], [Shuto, 1987]. In this study, dn was calculated as:
2
dn ¼ pffiffiffi Wd  102 ; ð10Þ
3D2
where 102 in the above equation adjusts a unit in Shuto's de¯nition of dn (No. cm) on a
1:100 scale because d, D, and W are in centimeters. In the experiments of Pasha and
Tanaka [2017] dn was set to three values, i.e.  180, 380, and 580 (No.cm).
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

According to the forest classi¯cation in terms of degree of damage by Shuto [1987], dn


180 No. cm represents areas where some trees on weak soil or at the forest fringe may
be damaged and soil around the trees may be scoured; dn 380 No.cm corresponds to
an area where soil in the forest may be scoured and damaged to some extent; and dn
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

580 No.cm represents an area where neither trees nor soil are damaged.

2.3. Finding the minimum water depth (y2) inside the length
of a hydraulic jump
Figure 1(a) shows the water pro¯le through vegetation. A previous literature review
found many studies for calculating the water depth after the hydraulic jump (y3 Þ
with reference to minimum water depth at the toe of hydraulic jump (y2 Þ [Chow,
1959; Chaudhry, 2008], but none determined the water depth at the toe of a
hydraulic jump with reference to the upstream water depths like h1 or h2 [Fig. 1(a)].
In this study, an attempt was made to ¯nd the y2 value with respect to h2 . The
results of the analytical method are compared with the actual experimental results of
Pasha and Tanaka [2017]. A continuity equation between the last row of vegetation
and the hydraulic jump can be written as:
h2 Vh  ¼ y2 Vy ; ð11Þ
where h2 and Vh are water depth and velocity in front of the last row of vegetation,
respectively, and y2 and Vy are water depth and velocity at the toe of the hydraulic
jump, respectively.  is the porosity.
Anjum et al. [2018] investigated the °ow properties and turbulence characteristics
in a circular vegetation patch and found that the stream-wise velocities are maximum
in the gap between two cylinders. In this study, the gap between the two consecutive
staggered cylinders of last row of vegetation (S ¼ D=2) was very small [Fig. 1(b) in this
study], so the stream-wise velocities were high within the gap. However, the °ow
decelerates behind the vegetation. Even in this situation, due to of inertia, the °ow can
keep its velocity for some distance (as shown in Fig. 5 in Anjum et al., 2018), and the
water surface gradient (inside the vegetation) can continue for some distance behind
the vegetation as well due to the \inertia e®ect" [as shown in Fig. 1(a) in this study].
The experimental results of Pasha and Tanaka [2017] show that, against constant
discharge conditions, the average velocities calculated by the equation of continuity
at the vegetation back are smaller (large water depth) as compared to the toe of

1950004-8
Critical Resistance of Vegetation

hydraulic jump (low water depth) without considering the porosity e®ect. The dif-
ference in sparse vegetation was 10–15%, in intermediate vegetation, the di®erence
varied from 15% to 30%, while in dense vegetation it was very large, i.e., 60–80%.
The estimated data shows that the denser the vegetation, the larger the di®erence
between the gap and the averaged velocity. However, due to secondary °ow currents
of induced turbulence inside the vegetation, the actual velocity (Vh Þ in front of the
last row vegetation (in between two cylinders) was much higher than the estimated
average velocity and the inertia e®ect behind the vegetation, thus considering
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

Vh ¼ Vy . Figure 4 also supports our assumption because a comparison of the ana-


lytical water depth \y 002 , which is calculated assuming that Vh ¼ Vy , and the exper-
imental water depth \y2 ", suggests a good agreement between analytical and
experimental results of water depth at the toe of hydraulic jump. Thus, Eq. (11)
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

becomes:
h 2  ¼ y2 : ð12Þ

2.4. Porosity
Figure 2 shows the de¯nition of porosity (). The maximum and minimum values of
porosity, which vary with vegetation conditions, are shown in Fig. 2(a) and Table 2.
However, the porosity value will be more accurate if the exact reduction width (rw)
in the wake of a cylinder is calculated, and it is represented by P , i.e. rwþ1 in
Fig. 2(b). The value of P depends upon the Froude number and L=d spacing. After
calculating the value of P , the porosity was determined.

2.4.1. Two-dimensional wake behind a single body


The two-dimensional wake behind single body was investigated by Schlichting
[1979]. Using the same approach, the width of the wake behind a cylinder against
all the values of G=d ratio was determined in this study. The equation of

(a) (b)

Fig. 2. De¯nition of porosity (Þ (a) maximum and minimum values, (b) wake formation behind cylinder
having a width of \P ".

1950004-9
G. A. Pasha & N. Tanaka

Table 2. Maximum and minimum


porosities with vegetation conditions

Type of vegetation G/d Porosity ()

Dense 0.25 20    60
Intermediate 1.09 52    76
Sparse 2.13 68    84
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

two-dimensional wake behind single body by Schlichting [1979] is given as:


pffiffiffiffiffi
b ¼ 10ðxCD dÞ1=2 ð13Þ
where b ¼ width of the wake, which is represented by rw in this study,  ¼
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

constant ¼ l=b; l ¼ mixing length de¯ned as l ¼ ky, (k ¼ von Karman's constant,


which has an approximate value of 0.41, and y is the half frontal projected width of a
cylinder, i.e. 0.5*diameter of cylinder), x ¼ distance from the center of a cylinder,
CD ¼ drag coe±cient (2.0, 1.3, 1.0 for dense, intermediate, and sparse, respectively
as suggested by Pasha and Tanaka [2017]), and d ¼ diameter of a cylinder.

3. Results
3.1. Validation of analytical method
The analytical method to ¯nd water depths was developed in Sec. 2 using the
momentum analysis approach. The mean values of drag coe±cients from the
experimental study by Pasha and Tanaka [2017] were used. The method was vali-
dated by comparing the calculated water depth at vegetation front and back with the
experimental results in Pasha and Tanaka [2017]. Figures 3(a) and 3(b) show the

(a)

Fig. 3. Correlation of analytical results and experimental results, (a) vegetation front, and (b) vegetation
back.

1950004-10
Critical Resistance of Vegetation
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

(b)

Fig. 3. (Continued )

correlation of analytical and experimental results of a backwater rise at vegetation


front (hfront Þ and back (hback Þ, respectively. It was revealed by Pasha and
Tanaka [2017] that the backwater rise decreased with either decreasing density or
thickness of vegetation. Figure 3(a) suggests a good agreement between analytical
and experimental results on backwater rise, although the analytical method pro-
duced a slightly larger value than the experiments at the vegetation back. In addi-
tion, in most of the cases, the di®erence between analytical and experimental
backwater rises for all the cases was less than 10%. The same approach to determine
the water level can be used for other vegetation and °ow conditions, but the correct
drag coe±cient must be known.

3.2. Wake behind a cylinder


After determining the width of the wake using a two-dimensional wake behind a
single body (Eq. 13) followed by porosity, the minimum water depth at the toe of the
hydraulic jump (y2 Þ was found for (i) dense, (ii) intermediate, and (iii) sparse den-
sities by using Eq. (12), and is shown in Fig. 4. The respective Froude number at y2 is
also shown in Fig. 5. Figure 5(a) suggested that the analytical Fr2 for dense vege-
tation is slighter higher for two initial °ow conditions (Fro ¼ 0:57, 0.62); however, it
is better validated for higher °ow conditions than the experimental values. Both the
experimental and analytical results suggest that vegetation thickness has little e®ect
on Fr2 . On average, dense vegetation transformed a subcritical °ow with Fro ¼ 0:69
to supercritical °ows with Fr2 ¼ 1:80 (experimental) and 1.94 (analytical).
Figure 5(b) shows the transformation of subcritical Fro to supercritical Fr2 for in-
termediate vegetation. On average, intermediate vegetation transformed a subcrit-
ical °ow with Fro ¼ 0:69 to a supercritical °ow with Fr2 ¼ 1:21 (experimental) and
1.13 (analytical). According to analytical results using sparse vegetation [Fig. 5(c)],

1950004-11
G. A. Pasha & N. Tanaka
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

(a)

(b)

(c)

Fig. 4. Correlation of analytical and experimental results of minimum water depth at toe of hydraulic
jump, y2 for (a) dense vegetation, (b) intermediate vegetation, (c) sparse vegetation, where dn has units of
No.cm.

1950004-12
Critical Resistance of Vegetation
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

(a)

(b)

(c)

Fig. 5. Transformation of subcritical °ow with initial Froude number, Fro to supercritical °ow with
Froude number Fr2 for (a) dense vegetation, (b) intermediate vegetation, and (c) sparse vegetation, where
ANA and EXP represents analytical and experimental data, respectively. dn has units of No.cm.

1950004-13
G. A. Pasha & N. Tanaka

the subcritical Fro remained in a subcritical state after passing through the vege-
tation, and Fr2 values were less than 1. However, in experiments, only sparse
vegetation with dn-380 transformed it into a supercritical °ow.

4. Discussion
In order to ¯nd the minimum water depth located in front of the formation of a
hydraulic jump, a porosity approach was utilized. In the current study, the e®ect of
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

°ow velocity was not included in the analytical approach to ¯nd the wake width;
therefore, a constant porosity value was used for every °ow condition. The drag
coe±cient (Cd ) changes with °ow conditions and is responsible for accurate deter-
mination of wake width.
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

When the approach of a two-dimensional wake behind a single body was used
[Schlichting, 1979], the analytical results matched the experimental results reported
in Pasha and Tanaka [2017] with good agreement, except for °ow with a very low
Froude number. From Figs. 4(a) and 4(b), it can be seen that the experimental and
analytical results are less di®erent, which validates the analytical results. Takemura
and Tanaka [2007] classi¯ed the G=d ratio of a colony-type emergent roughness
model according to instabilities of the vortex street. According to the ¯ndings of
Takemura and Tanaka [2007], the generation of a large Karman vortex street (LKV)
occurred for G=d < 0:4 while the primary Karman vortex street (PKV) behind
individual cylinders was generated for G=d > 1:8. Thus, based on the study of
Takemura and Tanaka [2007] three G=d ratios were selected to represent the three
di®erent °ow structures i.e. G=d < 0:4 (dense), 0:4 < G=d < 1:8 (intermediate), and
G=d > 1:8 (sparse). For dense and intermediate vegetation, the small vortices/eddies
were not prominent due to the large Karman vortex street °ow, and therefore they
did not a®ect the results. However, for sparse vegetation, small vortices/eddies were
generated that did a®ect the results. In Eq. (13) of the two-dimensional wake behind
a single body, the wake °ow width equation is basically for calculation of the wake
width of a single cylinder; therefore, due to LKV in dense and intermediate vege-
tation, the results are validated well. Moreover, the drag coe±cient for sparse veg-
etation was considered to be 1.0, which is slightly less for calculating the wake
reduction width.
Figure 5(a) shows that dense vegetation (G=d  0:25) transformed the subcritical
°ow (without vegetation) to a supercritical °ow (downstream of vegetation) with
the range of Froude numbers nearly 1.6–1.9 at minimum water depth. This results
in the formation of a breaking undular hydraulic jump for most of the cases and a
non-breaking undular hydraulic jump for a few cases with a low Froude number
°ow, whereas the intermediate vegetation (G=d  1:09) transformed the °ow to
critical °ow with a relatively low range of Froude numbers ranging from 1.1 to 1.2
[Fig. 5(b)]. This resulted in a non-breaking undular hydraulic jump. However,
according to the analytical results, except for few cases, sparse vegetation
(G=d  2:13) did not transform °ow to a critical °ow, and the range of Froude

1950004-14
Critical Resistance of Vegetation

numbers was from 0.85 to 0.98 [Fig. 5(c)]. Thus, in the current conditions, the
vegetation with G=d  1:0 acts as a critical condition. Decreasing the ratio i.e.
G=d < 1 transforms the °ow to a supercritical °ow. For estimating y2 accurately for
sparse conditions, Eq. (12) should be improved because the backwater rise is not
large, the water surface gradient is small, and the inertia e®ect, we assumed, may be
small in that case.
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

5. Conclusions
The critical resistance in terms of non-dimensional vegetation density (G=d) was
investigated. An analytical approach was developed to ¯nd the backwater rise and
downstream vegetation water depth, which were validated well by the laboratory
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

experiments. The analytical approach accurately determined the water depth in the
upstream vegetation, while on the downstream side, the water was slightly deeper
than in the experimental results. The two-dimensional wake behind a single body
was used to ¯nd the porosity of the last row of vegetation cylinders, which was
chosen to ¯nd the minimum water depth inside the hydraulic jump. In this study,
the dense vegetation (G=d  0:25) transformed the subcritical °ow (without vege-
tation) to a supercritical °ow (downstream of vegetation) with a range of Froude
numbers from nearly 1.6–1.9 at minimum water depth. Intermediate vegetation
(G=d  1:09) transformed the °ow to a critical °ow with a relatively low range of
Froude numbers ranging from 1.1 to 1.2. However, according to the analytical
results, except for a few cases, sparse vegetation (G=d  2:13) did not transform the
°ow to a critical °ow and the range of Froude numbers was from 0.85 to 0.98. Thus,
G=d  1:0 acted as the critical condition, below which the °ow will change from a
subcritical to a supercritical °ow.

Acknowledgments
This study was funded by a JSPS Grant-in-Aid for Scienti¯c Research
(No. 15H02987).

References
Anjum, N., Ghani, U., Pasha, G. A., Rashid, M. U., Latif, A. and Rana, M. Z. Y. [2018]
\Reynolds stress modeling of °ow characteristics in a vegetated rectangular open
channel," Arabian J. Sci. Eng. 43, 5551–5558.
Bokaian, A. and Geoola, F. [1984] \Wake-induced galloping of two interfering circular
cylinders," J. Fluid Mech. 146, 383–415.
Chaudhry, M. H. [2008] Open Channel Flow, 2nd edn. (Springer Publishing Co, New York).
Chow, V. T. [1959] Open Channel Hydraulics (McGraw-Hill Publishing Co., New York).
Cochard, R., Ranamukhaarachchi, S. L., Shivakoti, G. P., Shipin, O. V., Edwards, P. J. and
Seeland, K. T. [2008] \The 2004 tsunami in Aceh and Southern Thailand: A review on
coastal ecosystems, wave hazards and vulnerability, Perspectives in plant ecology," Evol.
Syst. 10, 3–40.

1950004-15
G. A. Pasha & N. Tanaka

Danielsen, F., Sorensen, M. K., Olwig, M. F., Selvam, V., Parish, F., Burgess, N. D., Hiraishi,
T., Karunagaran, V. M., Rasmussen, M. S., Hansen, L. B., Quarto, A. and Suryadiputra,
N. [2005] \The Asian tsunami: A protective role for coastal vegetation," Science 310, 643.
Dengler, L. and Preuss, J. [2003] \Mitigation lessons from the July 17, 1998, Papua New
Guinea tsunami," Pure Appl. Geophys. 160, 2001–2031.
Harada, K. and Imamura, F. [2000] \Experimental study on the resistance by mangrove under
the unsteady °ow," Proc. 1st Congress of the Asian and Paci¯c Coastal Engineering,
Dalian, pp. 975–984.
Harada, K. and Imamura, F. [2005] \E®ects of coastal forest on tsunami hazard mitigation–a
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

preliminary investigation, tsunamis: Case studies and recent development," in Satake, K.


(ed.), Advances in Natural and Technological Hazards Research (Springer), pp. 279–292.
Hiraishi, T. and Harada, K. [2003] \Greenbelt tsunami prevention in South-Paci¯c region,"
Rep. Port Airport Res. Inst. 42(2), 23.
Huang, Z., Wu, T. S., Chen, T. Z. and Sim, S. Y. [2013] \A possible mechanism of destruction
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

of coastal trees by tsunamis: A hydrodynamic study on e®ects of coastal steep hills,"


J. Hydro-Environment Res. 7, 113–123.
Iimura, K. and Tanaka, N. [2012] \Numerical simulation estimating the e®ects of tree density
distribution in coastal forest on tsunami mitigation," Ocean Eng. 54, 223–232.
Iimura, K. and Tanaka, N. [2013] \Dangerous zone formation behind ¯nite-length coastal
forest for tsunami mitigation," J. Earthq. Tsunami 7(4), 1350034.
Irtem, E., Gedik, N., Kabdasli, M. S. and Yasa, N. E. [2009] \Coastal forest e®ects on tsunami
run-up heights," Ocean Eng. 36, 313–320.
Kathiresan, K. and Rajendran, N. [2005] \Coastal mangrove forests mitigated tsunami,"
Estuar. Coast. Shelf Sci. 65(3), 601–606.
Li, Y., Du, W., Yu, Z., Tang, C., Wang, Y., Anim, D. O., Ni, L., Lau, J., Chew, S. A.
and Acharya, K. [2015] \Impact of °exible emergent vegetation on the °ow turbulence
and kinetic energy characteristics in a °ume experiment," J. Hydro-Environment Res. 9,
354–367.
Mascarenhas, A. and Jayakumar, S. [2008] \An environmental perspective of the post-
tsunami scenario along the coast of Tamil Nadu, India: Role of sand dunes and forests,"
J. Environ. Manage. 89, 24–34.
Nandasena, N. A. K., Tanaka, N. and Tanimoto, K. [2008] \Perspective of coastal vegetation
patches with topography variations for tsunami protection in 2D-numerical modeling,"
Annu J. Hydr. Eng. JSCE 52, 133–138.
Nandasena, N. A. K., Sasaki, Y. and Tanaka, N. [2012] \Modeling ¯eld observations of the
2011 Great East Japan tsunami: E±ciency of arti¯cial and natural structures on tsunami
mitigation," Coastal Eng. 67, 1–13.
Nateghi, R., Bricker, J. D., Guikema, S. D. and Bessho, A. [2016] \Statistical analysis of the
e®ectiveness of seawalls and coastal forests in mitigating tsunami impacts in Iwate and
Miyagi prefectures," PLoS ONE 11(8), e0158375.
Noarayanan, L., Murali, K. and Sundar, V. [2012] \Performance of °exible emergent
vegetation in staggered con¯guration as a mitigation measure for extreme coastal dis-
asters," Nat. Hazards 62, 531–550.
Osti, R., Tanaka, S. and Tokioka, T. [2009] \The importance of mangrove forest in tsunami
disaster mitigation," Disasters 33(2), 203–213.
Pasha, G. A. and Tanaka, N. [2016a] \E®ectiveness of ¯nite length inland forest in trapping
tsunami-borne wood debris," J. Earthq. Tsunami 10(4), 1650008.
Pasha, G. A. and Tanaka, N. [2016b] \Energy loss and drag in a steady °ow through emergent
vegetation," Proc. of 12th ICHE Conf., 6–10 November 2016, Tainan, Taiwan, pp. 1–6.

1950004-16
Critical Resistance of Vegetation

Pasha, G. A. and Tanaka, N. [2017] \Undular hydraulic jump formation and energy loss in a
°ow through emergent vegetation of varying thickness and density," Ocean Eng. 141,
308–325.
Pasha, G. A., Tanaka, N., Yagisawa, J. and Achmad, F. N. [2018] \Tsunami mitigation by
combination of coastal vegetation and a backward-facing step," Coastal Eng. J.,
doi: 10.1080/21664250.2018.1437014.
Reconstruction Agency. [2011] \Basic Guidelines for Reconstruction in Response to the
Great East Japan Earthquake." 47 pages. Available at http://www.reconstruction.go.jp/
english/pdf/Basic Guidelines for Reconstruction.pdf.
by WSPC on 04/08/19. Re-use and distribution is strictly not permitted, except for Open Access articles.

Schlichting, H. [1979] Boundary-Layer Theory (McGraw-Hill Publishing Co., New York).


Shuto, N. [1987] \The e®ectiveness and limit of tsunami control forests," Coastal Eng. Jpn.
30(1), 143–153.
Sturm, W. T. [2001] Open Channel Hydraulics (McGraw-Hill Publishing Co., New York).
Takemura, T. and Tanaka, N. [2007] \Flow structures and drag characteristics of a colony–
J. Earthquake and Tsunami Downloaded from www.worldscientific.com

type emergent roughness model mounted on a °at plate in uniform °ow," Fluid Dyn. Res.
39, 694–710.
Tanaka, N. [2009] \Vegetation bioshields for tsunami mitigation: Review of the e®ectiveness,
limitations, construction, and sustainable management," Landscape Ecol. Eng. 5, 71–79.
Tanaka, N. [2012] \E®ectiveness and limitations of coastal forest in large tsunami: Conditions
of Japanese pine trees on coastal sand dunes in tsunami caused by Great East Japan
Earthquake," J Jpn Soc Civil Eng Ser B1 (Hydraul. Eng.) 68(4), II 7–II 15.
Tanaka, N. and Ogino, T. [2017] \Comparison of reduction of tsunami °uid force and addi-
tional force due to impact and accumulating after collision of tsunami-produced driftwood
from a coastal forest with houses during the Great East Japan tsunami," Landscape Ecol.
Eng. 13, 287–304.
Tanaka N. and Onai A. [2017] \Mitigation of destructive °uid force on buildings due to
trapping of °oating debris by coastal forest during the Great East Japan tsunami,"
Landscape Ecol. Eng. (in press), doi: 10.1007/s11355-016-0308-4.
Tanaka, N., Sasaki, Y., Mowjood, M. I. M. and Jinadasa, K. B. S. N. [2007] \Coastal vege-
tation structures and their functions in tsunami protection: Experience of the recent
Indian Ocean tsunami," Landscape Ecol. Eng. 3, 33–45.
Tanaka, N., Yagisawa, J. and Yasuda, S. [2013] \Breaking pattern and critical breaking
condition of Japanese pine trees on coastal sand dunes in huge tsunami caused by Great
East Japan Earthquake," Nat. Hazards 65, 423–442.
Tanaka, N., Yasuda, S., Iimura, K. and Yagisawa, J. [2014] \Combined e®ects of coastal forest
and sea embankment on reducing the washout region of houses in the Great East Japan
tsunami," J. Hydro-Environment Res. 8, 270–280.
Thuy, N. B., Tanimoto, K., Tanaka, N., Harada, K. and Iimura, K. [2009] \E®ect of open gap
in coastal forest on tsunami run-up – investigations by experiment and numerical
simulation," Ocean Eng. 36, 1258–1269.
Yanagisawa, H., Koshimura, S., Goto, K., Miyagi, T., Imamura, F., Ruangrassamee, A. and
Tanavud, C. [2009] \The reduction e®ects of mangrove forest on a tsunami based on ¯eld
surveys at Pakarang Cape, Thailand and numerical analysis," Estuarine Coastal Shelf
Sci. 81, 27–37.
Yanagisawa, H., Koshimura, S., Miyagi, T. and Imamura, F. [2010] \Tsunami damage
reduction performance of a mangrove forest in Banda Aceh, Indonesia inferred from ¯eld
data and a numerical model," J. Geophys. Res. 115, C06032.
Yokojima, S., Kawahara, Y. and Yamamoto, T. [2015] \Impacts of vegetation con°guration
on °ow structure and resistance in a rectangular open channel," J. Hydro-Environment
Res. 9, 295–303.

1950004-17

You might also like