Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

Applied Ocean Research 21 (1999) 81–91

Wave interactions with double slotted barriers


Michael Isaacson*, John Baldwin, Sundarlingam Premasiri, Gang Yang
Department of Civil Engineering, University of British Columbia, Vancouver, B.C., Canada V6T 1Z4
Received for publication 4 December, 1998

Abstract
The present article outlines the numerical calculation of wave interactions with a pair of thin vertical slotted barriers extending from the
water surface to some distance above the seabed, and describes laboratory tests undertaken to assess the numerical model. The numerical
model is based on an eigenfunction expansion method and utilizes a boundary condition at the surface of each barrier which accounts for
energy dissipation within the barrier. Comparisons with experimental measurements of the transmission, reflection, and energy dissipation
coefficients for partially submerged slotted barriers show excellent agreement and indicate that the numerical method is able to adequately
account for the energy dissipation by the barriers. 䉷 1999 Elsevier Science Ltd. All rights reserved.
Keywords: Breakwaters; Coastal engineering; Coastal structures; Permeability; Wave reflection; Wave transmission; Waves

1. Introduction free surface to the seabed (e.g. Sollitt and Cross [19]; Hagi-
wara [5]; Bennet et al. [2]; Yu [26]). Most formulations
Breakwaters are widely used to provide economical assume the velocity through the porous medium is propor-
protection from waves in harbours and marinas. In certain tional to the pressure gradient, with a complex proportion-
situations, breakwaters in the form of thin, rigid, pile- ality constant that accounts for possible phase differences
supported vertical barriers which extend some distance between the velocity and pressure gradient. This description
down from the water surface have been used or considered. can be related to the physics of the flow within the structure
These have the advantages of allowing water circulation, on the basis of both frictional and inertial effects. Compar-
fish passage and sediment transport beneath the breakwater, isons with experimental measurements of transmission and
and may be relatively economical by providing protection reflection coefficients have been carried out for slotted
closer to the water surface where wave action is most barriers extending to the seabed by Hagiwara [5]; Kriebel
pronounced. Predictions of wave interactions with such [10]; and Bennet et al. [2]) and generally exhibit satisfactory
structures have been obtained previously by a number of agreement. The extension to the case of a partially
authors for the case of an impermeable barrier on the submerged barrier has been considered more recently by
basis of linear wave diffraction theory. Numerical solutions Isaacson et al. [7].
have been developed on the basis of the boundary element Permeable barriers have the advantage of reducing wave
method (Liu and Abbaspour [11]; Nakamura [16]) and the reflection on the upwave side of the barrier but in order to
eigenfunction expansion method (Losada et al. [12]; Abul- also reduce wave transmission to an acceptable level it is
Azm [1]). often necessary to use two vertical barriers in many practical
In some instances, a permeable barrier, such as a slotted applications. A number of authors have considered the
vertical barrier made from timber planks, may be preferred. problem of two or more barriers. These include several
For example, this may be selected in an effort to reduce studies on using multiple porous plates as wave absorbers,
unwanted wave reflections on the upwave side of the barrier. e.g. Twu and Lin [22,23] and Losada et al. [13]; experimen-
Thus, the prediction of wave interactions with a permeable tal studies of a double screen breakwater by Gardner et al.
or slotted thin vertical barrier is also of interest. A primary [4] and comparisons between theoretical and experimental
feature of such interactions is that wave energy is absorbed results for wave reflection and transmission by double verti-
within the structure. Several authors have considered such cal slotted barriers, e.g. Kondo [9] and Hagiwara [5]. All of
predictions for permeable structures that extend from the the above studies have been carried out for barriers extend-
ing from the free surface to the seabed, and little work has
* Corresponding author. Tel.: 001 604 822 6412; Fax: 001 604 822 7006; been reported on double barriers that are partially
e-mail: isaacson@apsc.ubc.ca submerged.
0141-1187/99/$ - see front matter 䉷 1999 Elsevier Science Ltd. All rights reserved.
PII: S0141-118 7(98)00039-X
82 M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91
p
⫺1; t is time, k is the wave number, and g is the gravita-
tional constant.
The fluid domain is subdivided into three regions by the
planes of the barriers, as shown in Fig. 1, and the two-
dimensional potential f in Eq. (1) is denoted f 1, f 2 and
f 3 in regions 1, 2 and 3 respectively. Along x ˆ ^l, the
pressure (and hence the velocity potential) and the horizon-
tal velocity are equated along the matching boundaries
within the fluid, and are applied to a suitable boundary
condition along the surface of each barrier. The conditions
along the matching boundaries are thus:
2f 1 2f 2
f1 ˆ f2 ; ˆ along x ˆ ⫺l for 0 ⱕ z ⱕ a;
Fig. 1. Definition sketch. 2x 2x
…3†
In the present article an eigenfunction method is devel-
oped for wave interactions with a pair of thin permeable 2f 2 2f 2
f2 ˆ f3 ; ˆ along x ˆ ⫹l for 0 ⱕ z ⱕ a;
barriers extending from the water surface to some distance 2x 2x
above the seabed. This is an extension of a method devel- …4†
oped by Isaacson et al. [7] for a single partially submerged where a ˆ d ⫺ h (see Fig. 1).
vertical permeable barrier. The method is used to develop a
numerical model and theoretical predictions are compared
2.2. Permeable boundary condition
with the results from laboratory tests undertaken to assess
the numerical model. The boundary condition along the permeable barriers
may be developed on the basis of the formulation of Sollitt
and Cross [19] (see also Sulisz [21]), and as adopted by Yu
2. Theoretical formulation [26] for a thin vertical barrier extending to the seabed. This
may be expressed as:
2.1. Governing equations
2f1 2f2
ˆ ˆ ⫺iG 0 …f2 ⫺ f1 †
A normally incident, regular, small amplitude wave train 2x 2x …5†
of height H and angular frequency v propagates in water of along x ˆ ⫺l for a ⱕ z ⱕ d;
constant depth d past two identical thin permeable vertical
barriers a distance 2l apart, as shown in Fig. 1. The barriers
2f2 2f3
extends downwards a distance h below the still water level. ˆ ˆ ⫺iG 0 …f3 ⫺ f2 †
A Cartesian coordinate system (x, z) is defined with x 2x 2x …6†
measured in the direction of wave propagation from a along x ˆ ⫹l for a ⱕ z ⱕ d;
point mid-way between the barriers, and z measured
upwards from the seabed. The fluid is assumed incompres- where G 0 ˆ G=b, b is the barrier thickness and G is a perme-
sible and inviscid, and the flow irrotational. The fluid ability parameter which is generally complex. Eqs. (5) and
motion can therefore be described by a velocity potential (6) correspond to the fluid velocity normal to the barrier
F which satisfies the Laplace equation within the fluid being proportional to the pressure difference across the
region. In addition, the wave height is assumed sufficiently barrier, with a complex constant of proportionality so that
small for linear wave theory to apply. Consequently, F is the real part of G corresponds to the resistance of the barrier
subject to the usual boundary conditions, linearized where and the imaginary part of G corresponds to the phase differ-
appropriate, at the seabed, free surface and far field (see, for ences between the velocity and the pressure because of
example, Sarpkaya and Isaacson [18]), and may thus be inertial effects. Various authors have related the permeabil-
expressed, using complex notation, in the form: ity parameter, G, to the physics of the flow within the barrier
in different ways. For example the resistance may be
F…x; z; t† ˆ Re‰Cf…x; z†exp…⫺ivt†Š; …1† expressed in terms of a friction coefficient (e.g. Sollitt and
Cross [19]), a drag coefficient (e.g. Hagiwara [5]) or a head
where
loss coefficient (e.g. Mei et al. [15]), while the inertial
  effects are generally expressed in terms of an added mass
igH 1
Cˆ⫺ : …2† coefficient (e.g. Sollitt and Cross [19]) or an effective orifice
2v cosh…kd†
length (e.g. Mei et al. [15]). In the present article, the
Also, Re[ ] denotes that the real part of the argument, i ˆ method of Sollitt and Cross [19] is followed and G is
M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91 83

X

f3 ˆ A4m cos…mm z†exp‰⫺mm …x ⫺ l†Š …x ⱖ ⫹l†: …11†
mˆ0

Here f w represents the incident wave potential which is


given as:
fw ˆ cosh…kz†exp…ikx†: …12†

Also, m m for m ⱖ 1 are the positive real roots of the follow-


ing equation, taken in ascending order:

v2 d
mm d tan…mm d† ˆ ⫺ for m ⱖ 1: …13†
Fig. 2. Geometry of slotted barrier. g
m 0 itself corresponds to the imaginary root of the above
expressed by: equation, such that m0 ˆ ⫺ik, with the wave number k
1 being given as the real root of the corresponding equation:
Gˆ ; …7†
f ⫺ is
v2 d
kd tanh…kd† ˆ : …14†
where e is the porosity of the barrier (defined as the fraction g
of area occupied by the slots, c=…w ⫹ c† see Fig. 2), f is a
friction coefficient and s is an inertia coefficient given by Thus, Eqs. (9)–(11) each represent the incident wave train
  combined with a superposition of a propagating mode
1⫺1 (m ˆ 0) and a series of evanescent modes (m ⱖ 1) which
s ˆ 1 ⫹ Cm : …8†
1 decay with distance away from the barrier. They satisfy all
In Eq. (8), Cm is an added mass coefficient. It is noted that the relevant boundary conditions, except that the conditions
although the barrier is taken to have zero thickness with of pressure continuity along the matching boundary and the
respect to the wave diffraction problem under consideration, boundary condition at the barrier surface are still needed to
the barrier width is considered non-zero with respect to the determine the coefficients A1m, A2m, A3m, and A4m.
flow within the barrier. The friction coefficient, f, comes For 0 ⱕ z ⱕ a, the matching conditions expressed in Eqs.
from a linearization of the velocity squared term associated (3) and (4) give rise to the following set of equations for A1m,
with the head loss across the barrier. In the original formu- A2m, A3m, and A4m:
lation of Sollitt and Cross [19] f is calculated implicitly X
∞ X

using the Lorentz principle of equivalent work so that the A1m cos…mm z† ⫺ A2m cos…mm z†
nonlinear effects of wave steepness are retained. This mˆ0 mˆ0
requires an iterative procedure, and in the present article X

the formulation of Yu [26] is followed such that f is treated ⫺ A3m cos…mm z†exp…⫺2mm l†
simply as a constant which is assumed to be known. mˆ0

2.3. Eigenfunction expansion ˆ ⫺cos…m0 z†exp…m0 l†; …15†

Expressions for f 1, f 2 and f 3 which satisfy the seabed, X


∞ X

free surface, and radiation conditions, as well as the above A1m mm cos…mm z† ⫺ A2m mm cos…mm z†
conditions along x ˆ ^l, may be developed in terms of four mˆ0 mˆ0
sets of coefficients A1m, A2m, A3m, and A4m which are initially X

unknown: ⫺ A3m mm cos…mm z†exp…⫺2mm l†
mˆ0
X

f1 ˆ fw ⫹ A1m cos…mm z†exp‰mm …x ⫹ l†Š …x ⱕ ⫺l†; …9†
ˆ m0 cos…m0 z†exp…m0 l†; …16†
mˆ0

X
∞ X
∞ X

f2 ˆ A2m cos…mm z†exp‰⫺mm …x ⫹ l†Š A2m cos…mm z†exp…⫺2mm l† ⫹ A3 cos…mm z†
mˆ0 mˆ0 mˆ0

X

…10† X

⫹ A2m cos…mm z†exp‰mm …x ⫺ l†Š ⫺ A4m cos…mm z†
mˆ0 mˆ0

…⫺l ⱕ x ⱕ ⫹l†; ˆ 0; …17†


84 M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91

X
∞ X
∞ Aim. This gives rise to a matrix equation for A1m, A2m, A3m
A2m mm cos…mm z†exp…⫺2mm l† ⫺ A3m mm cos…mm z† and A4m:
mˆ0 mˆ0 2 ∞ 3
X …mn† X ∞
…mn†
X

…mn†
X

…mn†
X

6 C11 C12 C13 C14 7
⫺ A4m mm cos…mm z† 6 mˆ0 7
6 mˆ0 mˆ0 mˆ0 78 9
6 7
mˆ0 6X ∞ X
∞ X
∞ X∞ 7> A1m >
6 …mn† …mn† …mn† …mn† 7>> >
>
6 C24 7
ˆ 0; 6 C C22 C23 7>> >
>
<
7 A2m =
21
6 mˆ0
…18† 6 mˆ0 mˆ0 mˆ0
7
6X 7>
6 ∞ X
∞ X
∞ X∞ 7> A3m >
…mn† 7> >
>
6 …mn† …mn† …mn†
And along the barrier’s surface the boundary conditions 6 C C32 C33 C34 7>>
:
>
>
;
6 mˆ0 31 7
6 mˆ0 mˆ0 mˆ0 7 A4m
(expressed by Eqs. (5) and (6)), give rise to the following 6 ∞ 7
equations for a ⱕ z ⱕ d: 6X X
∞ X
∞ X∞ 7
4 …mn† …mn† …mn† …mn† 5
C41 C42 C43 C44
X
∞ X
∞ mˆ0 mˆ0 mˆ0 mˆ0
A1m cos…mm z†…mm d ⫺ iG 0 d† ⫹ iG 0 d A2m cos…mm z† 8 9
mˆ0 mˆ0 >
> b1n >
>
>
> >
>
X
∞ >
< b2n >
=
⫹ iG 0 d A3m cos…mm z†exp…⫺2mm l† ˆ ;
>
> b3n >
>
mˆ0 >
> >
>
>
: >
;
ˆ …m0 d ⫹ iG 0 d†cos…m0 z†exp…m0 l†; b4n
…19† …23†
where
X
∞ X

iG 0 d A1m cos…mm z† ⫹ A2m cos…mm z†…mm d ⫺ iG 0 d† …mn†
C11 ˆ fmn …0; a† ⫹ …mm d ⫺ iG 0 d†fmn …a; d†; …24†
mˆ0 mˆ0
…mn†
X
∞ C12 ˆ ⫺fmn …0; a† ⫹ iG 0 dfmn …a; d†; …25†
0
⫺ A3m cos…mm z†exp…⫺2mm l†…mm d ⫹ iG d†
…mn†
mˆ0
C13 ˆ exp…⫺2mm l†‰⫺fmn …0; a† ⫹ iG 0 dfmn …a; d†; …26†
ˆ ⫺iG 0 d cos…m0 z†exp…m0 l†;
…mn†
C21 ˆ mm dfmn …0; a† ⫹ iG 0 dfmn …a; d†; …27†
…20†
…mn†
X
∞ C22 ˆ mm dfmn …0; a† ⫹ …mm d ⫺ iG 0 d†fmn …a; d†; …28†
0
A2m cos…mm z†exp…⫺2mm l†…mm d ⫹ iG d†
…mn†
mˆ0 C23 ˆ exp…⫺2mm l†‰⫺mm dfmn …0; a† ⫺ …mm d
X
∞ X

⫺ A3m cos…mm z†…mm d ⫺ iG 0 d† ⫺ iG 0 d A4m cos…mm z† ⫹ iG 0 d†fmn …a; d†Š; …29†
mˆ0 mˆ0
…mn†
C32 ˆ exp…⫺2mm l†‰fmn …0; a† ⫹ …mm d ⫹ iG 0 d†fmn …a; d†Š;
ˆ 0;
…30†
…21†
…mn†
X
∞ X
∞ C33 ˆ fmn …0; a† ⫺ …mm d ⫺ iG 0 d†fmn …a; d†; …31†
0 0
iG d A2m cos…mm z†exp…⫺2mm l† ⫹ iG d A3m cos…mm z†
…mn†
mˆ0 mˆ0 C34 ˆ ⫺fmn …0; a† ⫺ iG 0 dfmn …a; d†; …32†
X

…mn†
⫹ A4m cos…mm z†…mm d ⫺ iG 0 d† C42 ˆ exp…⫺2mm l†‰mm dfmn …0; a† ⫹ iG 0 d†fmn …a; d†Š; …33†
mˆ0
…mn†
C43 ˆ ⫺mm dfmn …0; a† ⫹ iG 0 dfmn …a; d†; …34†
ˆ 0;
…22† …mn†
C44 ˆ ⫺mm dfmn …0; a† ⫹ …mm d ⫺ iG 0 d†fmn …a; d†; …35†
Eqs. (15)–(22) are integrated in an appropriate way and then
…mn† …mn† …mn†
added to develop a suitable matrix equation for Aim Thus, C14 ˆ C24 ˆ C31 ˆ C41
9mn†
ˆ 0; …36†
each equation is first multiplied by cos(m nz), then integrated
with respect to z over the appropriate domain of z (i.e., from b1n ˆ exp…m0 l†‰⫺f0n …0; a† ⫹ …m0 d ⫹ iG 0 d†f0n …a; d†Š; …37†
z ˆ 0 to a, or from z ˆ a to d), and each pair of resulting
equations is then added to obtain four sets of equations for b2n ˆ exp…m0 l†‰m0 df0n …0; a† ⫺ iG 0 df0n …a; d†Š; …38†
M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91 85

Finally, the runup R on the upwave side of the first


barrier, and the maximum horizontal force per unit width
on each barrier, F can be obtained as:

H 1 X∞

R ˆ exp…m0 l† ⫹ A1m cos…mm d† …44†
2 cosh…kd† mˆ0

rgH X X∞
F1 ˆ A1m dm ⫺ A2m dm
2 cosh…kd† mˆ0 mˆ0

X


⫺ A3m dm exp…⫺2mm l† ⫹ d0 exp…m0 l† …45†

mˆ0


Fig. 3. General view of the wave flume showing the barrier setup. rgH X X∞
F2 ˆ A2m dm exp…⫺2mm l† ⫹ A3m dm
2 cosh…kd† mˆ0 mˆ0
b3n ˆ b4n ; …39†
X


⫺ A4m dm ;
Zb
mˆ0
fmn …a; b† ˆ cos…mm z†cos…mn z†dz
a …46†
  where
1 sin…mz † sin…m⫹ z† b
ˆ ⫹ for m 苷 n
2 m m⫹ a sin…mm d† ⫺ sin…mm a†
dm ˆ : …47†
mm
1
and ‰2mm z ⫹ sin…2mm z†Šba for m ˆ n:
4mm
…40† 3. Description of experiments
Also m⫺ ˆ mm ⫺ mn , and m⫹ ˆ mm ⫹ mn .
Experiments were carried out in the wave flume of the
2.4. Numerical solution Hydraulics Laboratory of the Department of Civil Engineer-
ing at the University of British Columbia. The flume is 20 m
In a numerical solution to the problem, Eq. (23) is trun- long and 0.62 m wide. Waves are generated by a single
cated to a finite number of terms N, and thus becomes a paddle wave actuator located at the upwave end. An artifi-
complex matrix equation of rank 4N which can be solved cial beach which is covered by a mat of synthetic hair is
for the first N unknown values of each set of coefficients A1m, located at the downstream end of the flume in order to
A2m, A3m and A4m. Once these have been calculated, the minimize wave reflection. The vertical barriers to be tested
various quantities of engineering interest may readily be were placed in the test section 10 m from the wave genera-
obtained. The method is described in more detail by Yang tor. Fig. 3 shows a general view of the experimental setup.
[25]. If required, it may readily be extended in the usual way The permeable wave barriers are constructed of vertical
to the cases of oblique waves and/or irregular waves (e.g. panels of width w ˆ 2.0 cm, and thickness b ˆ 1.3 cm,
Dalrymple et al. [3], Losada et al. [12,14]). such that the porosity of the barriers can be varied by chan-
The (real) transmission and reflection coefficients, ging the dimensions of the slots between the panel members.
denoted Kt and Kr respectively, are defined as the appropri- Tests were carried out with a constant water depth of
ate ratios of wave heights: Kt ˆ Ht =H and Kr ˆ Hr =H, where 0.45 m and with generator motions corresponding to regular
Ht and Hr are the transmitted and reflected wave heights wave trains with five different wave periods
respectively. These are given in terms of A1m and A4m by: (T ˆ 0:6; 0:8; 1:0; 1:2 and 1.4 s), and specified wave heights
Kt ˆ jA40 j; …41† corresponding to a constant wave steepness, H=L ˆ 0:07.
Pairs of half and fully immersed barriers (h=d ˆ 0:5 and
Kr ˆ jA10 j: …42† 1.0 respectively) with porosities of 0, 5 and 10% and
spacings of l=d ˆ 0:22; 0:55 and 1.10 were tested.
From considerations of energy conservation, these are In each test, the water surface elevation was measured
related to the energy dissipation coefficient Ke: upwave and downwave from the wave barriers using twin
wire capacitance wave probes. Three probes were used on
Kr2 ⫹ Kt2 ⫹ Ke ˆ 1; …43†
the upwave side of the barriers at distances of 1.0, 1.2 and
where Ke is the proportion of the incident wave energy flux 1.5 m from the upwave barrier, and two probes were used on
that is dissipated by the barrier. the downwave side at distances of 1.0 and 1.2 m from the
86 M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91

Fig. 4. Comparison of predicted hydrodynamic coefficients with results


from Hagiwara [5] as functions of spacing, l h, for h=d ˆ 1:0, e ˆ 15% and
kh ˆ 1:5: (a) transmission coefficient, (b) reflection coefficient. —, present
method; and ⫹, Hagiwara [5].

downwave barrier. These probe positions were selected by


considering the evanescent wave modes from the barrier,
the wave paddle location and the method of reflection analy-
sis. Wave generation and data acquisition were controlled
by a VAX workstation, using the GEDAP (General Experi-
mental control, Data acquisition and Analysis Package)
software package developed at the Coastal Engineering
Laboratory of the National Research Council, Canada.
The wave probes were sampled at a rate of 20 Hz for a
duration of 14 s for each test run. Fig. 5. Comparison between measured and predicted hydrodynamic coeffi-
The wave records obtained from the probes were used to cients as functions of kh, for h=d ˆ 0:5, e ˆ 5% and l=h ˆ 0:44: (a) trans-
estimate the reflection and transmission coefficients for each mission coefficient, (b) reflection coefficient and (c) energy dissipation
test. The reflection and transmission coefficient were esti- coefficient. O, measured; —, predicted.
mated using a least squares method applied to simultaneous
measurements of the water surface elevation upwave and limiting case of two permeable barriers resting on the
downwave of the barriers (e.g. Isaacson [6]). The incident seabed (h/d ˆ 1.0). Fig. 4 shows a comparison of the
wave height was obtained from measurements without the predicted transmission and reflection coefficients with the
barriers in place. For a further description of the experi- numerical results of Hagiwara [5] as functions of the dimen-
ments, see Premasiri [17]. sionless spacing between the barriers l /h, for a relative
depth kd ˆ 1:5 and a porosity e ˆ 5%. A reasonably close
fit between the transmission and reflection coefficients
4. Results and discussion predicted by the two methods is obtained with a friction
coefficient of f ˆ 0:5 and an added mass coefficient of
4.1. Comparison with previous predictions Cm ˆ 0:18, despite minor differences in the two formula-
tions. Hagiwara [5] relates the resistance component of the
The numerical method was first developed for a single permeability parameter to a drag coefficient rather than the
barrier and was validated by comparison with previous friction coefficient used here and the resistance term is line-
predictions for the limiting cases of an impermeable barrier arized in a different way. Also Hagiwara [5] relates the
and a permeable barrier extending down to the seabed. This added mass to an effective orifice length incorrectly so
was described by Isaacson et al. [7]. In the present article the that the added mass Cm ˆ 2:19 that he quotes in fact corre-
method is extended to the case of two vertical barriers. sponds to Cm ⬇ e. Note that for a comparison to the experi-
Results for the general case of two partially submerged mental results of Hagiwara [5] for the same case, a better fit
permeable barriers are not available and the method is is obtained with f ˆ 2:0 and Cm ˆ 0:18.
validated by comparison with previous predictions for the The number of terms used in the eigenfunction expansion
M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91 87

Fig. 6. Comparison between measured and predicted hydrodynamic coeffi-


cients as functions of kh, for h=d ˆ 0:5, e ˆ 5% and l=h ˆ 1:1: (a) trans-
mission coefficient, (b) reflection coefficient and (c) energy dissipation
coefficient. O, measured; —, predicted.

Fig. 7. Comparison between measured and predicted hydrodynamic coeffi-


cients as functions of kh, for h=d ˆ 0:5, e ˆ 5% and l=h ˆ 2:2: (a) trans-
has been taken as N ˆ 100. This was found to give accurate mission coefficient, (b) reflection coefficient and (c) energy dissipation
results over the range of values presented here. coefficient. O, measured; —, predicted.

4.2. Comparison with experiments


larger relative spacings, peaks in the transmission and
In order to assess further the theoretical model, numerical reflection coefficients occur when the relative draft
predictions were compared with the results of experiments. kh ˆ np=…2l=h†, corresponding to resonant excitation of
Figs. 5, 6 and 7 show a comparison of the measured and partial standing waves between the barriers. These were
predicted transmission, reflection and energy dissipation observed experimentally and somewhat surprisingly corre-
coefficients as functions of kh for a relative draft spond to a reduction in the energy dissipation coefficient.
h=d ˆ 0:5, a porosity e ˆ 5% and for barrier spacings l=h ˆ The most notable difference between the numerical and
0:44; 1:10 and 2.20 respectively. The figures follow experimental results occurs in the transmission coefficient:
expected trends: the transmission coefficient decreases the measured transmission coefficients are consistently
with increasing kh, while the reflection coefficient follows lower than the predicted transmission coefficients at lower
the opposite trend, and the energy dissipation coefficient Ke relative drafts for all three barrier spacings. This is reflected
is non-zero accounting for the energy loss across the barrier. in higher energy dissipation coefficients for the experimen-
In general, the numerical model is able to adequately repro- tal results and is caused by additional energy losses asso-
duce the most important features of the experimental results, ciated with vortex formation around the bottom edges of the
including the energy dissipation through the slotted barrier, barriers. Such losses are generally difficult to estimate
and the measured and predicted hydrodynamic coefficients because of scatter in the experimental results, but are seen
compare reasonably well for all three barrier spacings. For to be as high as 30%. This is considerably higher than
88 M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91

energy losses for single barriers which were previously esti-


mated to be between 10% and 20% at lower relative drafts
for permeable [7] and impermeable barriers [8,20]. In the
case of the single barrier the energy losses because of vortex
formation are generally small when the transmission coeffi-
cient is low and thus are not likely of much importance for
practical applications [7]. In the case of the double barrier,
however, the energy losses because of vortex formation are
higher because of the additional barrier and could be signif-
icant within the range of some practical applications. Preli-
minary efforts to include vortex formation in a numerical
model are described by Nakamura [16].
In order to apply the numerical model, suitable values of
the friction and added mass coefficients are needed. It is
clear that the values of f and Cm, which are empirical by
definition, depend on the specific conditions of laboratory
tests and on the detailed geometry of the slots. An extensive
investigation of the variation of these coefficients over a full
range of experimental conditions and with slot geometry
was not carried out, and is beyond the scope of the present
study. For example, their values may differ somewhat for
different slot geometries. However, it is expected that the
coefficients will be reasonably constant over a wide range of
conditions, and the purpose of the laboratory tests carried
out here is to assess the numerical model on the basis of
constant coefficients. Thus, in the present study these coeffi-
cients were estimated from previous results with a single
barrier [7] on the basis of a best fit between the measured
and predicted values of the transmission, reflection and
energy dissipation coefficients. This procedure gives a fric-
tion coefficient f ˆ 2:0 and an added mass coefficient
Cm ˆ 0, which agrees with the values used by Yu [26] for
Fig. 8. Comparison between measured and predicted hydrodynamic coeffi-
cients as functions of dimensionless spacing, l /h, for h=d ˆ 0:5, e ˆ 5% rubblemound breakwaters extending down to the seabed,
and kh ˆ 0:95: (a) transmission coefficient, (b) reflection coefficient and (c) with the values of Cm used by Twu and Lin [22,23] and
energy dissipation coefficient. O, measured; —, predicted. Losada et al. [13] for multiple slotted barriers extending
down to the seabed, and with the values of Cm measured
by Urashima et al. [24] for a single slotted barrier extending
down to the seabed.
Fig. 8 gives a further comparison between the predictions
of the numerical model and the experimental results, and
shows the transmission, reflection and energy dissipation
coefficients as functions of the relative spacing between
the barriers for h=d ˆ 0:5 and for kh ˆ 0:95. Overall agree-
ment is seen to be satisfactory. It appears that the measured
variation of the hydrodynamic coefficients with spacing is
less than predicted by the numerical model. This effect was
also observed by Hagiwara [5] for the fully submerged
barrier.
In the design of a slotted breakwater, the choice of poros-
ity is particularly important. This is illustrated in Fig. 9
which shows the measured and predicted transmission and
reflection coefficients as functions of porosity for the half
submerged barriers at a relative draft kh ˆ 0:95 and a rela-
Fig. 9. Comparison between measured and predicted transmission and tive spacing l=h ˆ 1:1. There is some scatter in the experi-
reflection coefficients as functions of porositye , for h=d ˆ 0:5, l=h ˆ 1:1 mental results but again the overall agreement is
and kh ˆ 0:95. O, measured; —, predicted. satisfactory.
M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91 89

Fig. 10. Comparison of hydrodynamic coefficients for single and double barriers as functions of kh, for h=d ˆ 0:5, l=h ˆ 1:1 and e ˆ 5%: (a) transmission
coefficient, (b) reflection coefficient and (c) energy dissipation coefficient. —, single barrier; – – – double barrier.

4.3. Comparison with single barrier influence of barrier spacing can also be seen more clearly:
the performance of the double barrier is surprisingly insen-
It is also of interest to compare the performance of a sitive to the spacing between the barriers and is influenced
double barrier with that of a single barrier. Fig. 10 shows only by excitation of standing waves between the barriers.
a comparison of the transmission, reflection and energy Fig. 11 shows a comparison between the maximum hori-
dissipation coefficients of single and double barriers as func- zontal force on both the upwave and downwave barriers and
tions of kh, for a relative draft h=d ˆ 0:5, a porosity e ˆ 5%, the maximum horizontal force on a single barrier for a
and for double barrier spacings l=h ˆ 0:44; 1:10 and 2.20. barrier spacing l=h ˆ 0:44 and for h=d ˆ 0:5. For the
As expected, the addition of the second barrier has very little same degree of wave protection the wave force on the
effect on the reflection coefficient but gives a noticeable upwave barrier is less than the force on an equivalent single
decrease in the transmission coefficient, because of barrier. As expected the force on the downwave barrier is
increased energy dissipation by the double barrier. The considerably smaller.
90 M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91

Fig. 11. Comparison between the dimensionless maximum horizontal force on a single and double barrier as a function of kh, for h=d ˆ 0:5, l=h ˆ 0:44 and
e ˆ 5%. O, single barrier; – – – upwave and downwave barriers for the double barrier.

5. Conclusions [3] Dalrymple RA, Losada MA, Martin PA. Reflection and transmission
from porous structures under oblique wave attack. J. Fluid Mech.
1991;224:625–644.
The present article describes the numerical prediction of
[4] Gardner JD, Townend IH, Fleming CA. The design of a slotted verti-
wave interactions with a pair of thin slotted vertical barriers cal screen breakwater. Proc. 20th Coastal Eng. Conf., ASCE, Taipei,
extending from the water surface to some distance above the 1986, pp. 1881–1893.
seabed. The approach used is based on an eigenfunction [5] Hagiwara K. Analysis of upright structure for wave dissipation using
expansion method and utilizes a boundary condition at the integral equation. Proc. 19th Coastal Eng. Conf., ASCE, Houston,
1984, pp. 2810–2826.
surface of each permeable barrier which accounts for energy
[6] Isaacson M. Measurement of regular wave reflection. J. Waterway,
dissipation within the barrier. Expressions are developed for Port, Coastal and Ocean Eng. 1991;117(6):553–569.
parameters of engineering interest, including the transmis- [7] Isaacson M, Premasiri S, Yang G. Wave interactions with a vertical
sion and reflection coefficients, the wave runup, and the slotted barrier. J. Waterway, Port, Coastal and Ocean Eng. 1997,
maximum horizontal force on the barrier. submitted for publication.
[8] Knott GF, Mackley MR. On eddy motions near plates and ducts
Comparisons were carried out with previous numerical
induced by water waves and periodic flows. Phil. Trans. Roy. Soc.
studies for a permeable barrier extending down to the A 1980;294:599–623.
seabed, and close agreement was obtained [9] Kondo H. Analysis of breakwaters having two porous walls. Proc.
Laboratory tests were carried out to provide an assess- Coastal Structures ‘79, ASCE, Arlington, 1979, pp. 939–952.
ment of the numerical model. A comparison of correspond- [10] Kriebel DL. Vertical wave barriers: wave transmission and wave
forces. Proc. 23rd Coastal Eng. Conf., ASCE, Venice, 1992, pp.
ing numerical predictions of the transmission, reflection and
1313–1326.
energy dissipation coefficients with experimental results is [11] Liu PL-F, Abbaspour M. Wave scattering by a rigid thin barrier. J.
given. The agreement is generally satisfactory and indicates Waterway, Port, Coastal and Ocean Eng. Div. ASCE
that the present numerical method is able to adequately 1982;108(4):479–491.
account for the energy dissipation by the slotted break- [12] Losada IJ, Losada MA, Roldan AJ. Propagation of oblique incident
waves past rigid vertical thin barriers. Appl. Ocean Res.
waters, provided that the relevant empirical coefficients
1992;14:191–199.
are suitably chosen. The energy dissipation within the [13] Losada IJ, Losada MA, Baquerizo A. An analytical method to eval-
slotted barrier is related to friction and added mass coeffi- uate the efficiency of porous screens as wave dampers. Appl. Ocean
cients which have been estimated by fitting with experimen- Res. 1993;15:207–215.
tal results. A comparison with previous results for a single [14] Losada IJ, Losada MA, Losada R. Wave spectrum scattering by verti-
cal thin barrier. Appl. Ocean Res. 1994;16:123–128.
barrier shows that both the transmission coefficient and the
[15] Mei CC, Liu PLF, Ippen AT. Quadratic loss and scattering of long
maximum horizontal force on the upwave barrier are much waves. J. Waterway, Harbor and Coastal Eng. Div., ASCE
lower for a double barrier, while the reflection coefficient is 1974;100:217–239.
about the same. [16] Nakamura T. Numerical modeling of vortex formation around a large
angular body in waves. Proc., 2nd Int. Offshore and Polar Engrg.
Conf., ISOPE, San Francisco, 1992, vol. 3, pp. 217–224.
[17] Premasiri S. Experimental study of wave interactions with slotted
barriers. M.A.Sc. Thesis, Deptartment of Civil Engineering, Univer-
References sity of British Columbia, Vancouver, Canada, 1997.
[18] Sarpkaya T, Isaacson M. Mechanics of wave forces on offshore struc-
[1] Abul-Azm RG. Water diffraction through submerged breakwaters. J. tures, New York: Van Nostrand Reinhold, 1981.
Waterway, Port, Coastal and Ocean Eng. 1993;119(6):587–605. [19] Sollitt CK, Cross RH. Wave transmission through permeable break-
[2] Bennet GS, McIver P, Smallman JV. A mathematical model of a waters. Proc. 13th Coastal Eng. Conf., ASCE, Vancouver, 1972, pp.
slotted wavescreen breakwater. Coastal Engineering 1993;18:231– 1827–1846.
249. [20] Stiassnie M, Naheer E, Boguslavsky I. Energy losses due to vortex
M. Isaacson et al. / Applied Ocean Research 21 (1999) 81–91 91

shedding from the lower edge of a vertical plate attacked by surface [24] Urashima S, Ishizuka K, Kondo H. Energy dissipation and wave force
waves. Proc. Roy. Soc. Lon. A 1984;396:131–142. at slotted wall. Proc. 20th Coastal Engr. Conf., ASCE, Taipei, 1986,
[21] Sulisz W. Wave reflection and transmission at permeable breakwaters pp. 2344–2352.
of arbitrary cross section. Coastal Engineering 1985;9:371–386. [25] Yang G. Numerical model of wave effects on permeable vertical
[22] Twu SW, Lin DT. On highly effective wave absorber. Coastal Engi- barriers above the seabed. M.A.Sc. Thesis, Deptartment. of Civil
neering 1990;15:389–405. Engineering, University of British Columbia, Vancouver, Canada,
[23] Twu SW, Lin DT. Wave reflection by a number of thin porous plates 1997.
fixed in a semi-infinitely long flume. Proc. 22nd Coastal Eng. Conf., [26] Yu X-P. Diffraction of water waves by porous breakwaters. J. Water-
ASCE, Delft, 1990, pp. 1046–1059. way, Port, Coastal and Ocean Eng. 1995;121:275–282.

You might also like