Você está na página 1de 11

Water Science and Engineering, 2014, 7(4): 457-467

doi:10.3882/j.issn.1674-2370.2014.04.010
http://www.waterjournal.cn
e-mail: wse2008@vip.163.com

Design formulas of transmission coefficients for


permeable breakwaters
Su-xiang ZHANG1, 2, Xi LI*3
1. Key Laboratory of Meteorological Disaster of Ministry of Education, Nanjing University of
Information Science and Technology, Nanjing 210044, P. R. China
2. Department of Civil Engineering and Applied Mechanics, McGill University,
Montreal, Quebec H3A 2K6, Canada
3. College of Harbor, Coastal and Offshore Engineering, Hohai University, Nanjing 210098, P. R. China

Abstract: New empirical formulas of the transmission coefficient for permeable breakwaters
were suggested based on available experimental data regarding the low-crest structure (LCS),
including the permeable rubble mound breakwater and pile-type breakwater. The rationality of the
present formulas was verified by their comparison with existing empirical and analytical formulas.
Numerical flume results were obtained by solving the modified Boussinessq-type wave equations
(MBEs), and a new expression relating the friction coefficient α to the relative submerged depth
Rt H s was also derived. Comparative analysis shows that the results of the present formulas agree
with the numerical flume results as well as available experimental data, and the present formulas
are superior to the existing empirical and analytical expressions in estimating the transmission
coefficient. The present formulas can provide references for estimation of the transmission
coefficient in engineering practice.
Key words: permeable breakwater; transmission coefficient; wave reflection; nonlinear wave;
Boussinesq wave model; low-crest structure

1 Introduction
The main characteristic of a wave-permeable breakwater is that a wave can move through
the breakwater without changing its profile, while wave diffraction is not dependent on
whether the structure is permeable or not. Wave overtopping and transmission are the main
modes of energy transportation for waves moving through the permeable breakwater, with
some energy dissipated by wave breaking on the seaward side or by turbulence inside the
breakwater, and some energy reflected in front of the structure. Thus, a detailed description of
nearshore hydrodynamics is of engineering value in examining wave and structure interaction.
Although the nonlinear wave theory has been widely developed, it is still difficult to apply it
directly to engineering design because of the inconvenience in estimating the transmission
coefficient. Linear theories and empirical equations have a practical advantage in analysis of
—————————————
This work was supported by the Key Project in the National Science and Technology Pillar Program for the
Twelfth Five-Year Plan Period (Grant No. 2012BAB03B01), and the Jiangsu Provincial Post-Doctoral
Support Plan (Grant No. 20100197).
*Corresponding author (e-mail: xili@hhu.edu.cn)
Received Feb. 15, 2013; accepted Sep. 1, 2014
wave transformation, including an overall description of phenomena like wave breaking,
run-up, and overtopping. Permeable breakwaters, especially the pile type, have been widely
used in coastal protection, wave reduction, and ship berthing. It is useful for engineering
design to predict the transmitted wave height and transmission coefficient with a simple and
convenient empirical formula. In this study, the transmission coefficient of permeable
breakwaters was analyzed based on available experimental data regarding the low-crest
structure (LCS) (van der Meer et al. 2005; Chen et al. 2008; Li and Xie 2008; Peng et al.
2009), including an emergent porous rubble mound breakwater and a pile-type breakwater.
Based on available data (Madsen 1983; Luth et al. 1994; Christou et al. 2008; van der Meer
et al. 2005), two other permeable structures were also investigated: a porous rubble mound
breakwater before a rigid wall and a submerged breakwater. For the purpose of design
convenience, new empirical formulas of the transmission coefficient were suggested, and the
rationality was verified by their comparison with existing formulas, e.g., the analytical formula
of Wiegel (1961) and the empirical formulas of van der Meer and Daemen (1994) and
d’Angremond et al. (1996). Then, numerical flume results obtained by solving the modified
Boussinesq-type wave equations (MBEs) for pile-type breakwaters were analyzed and
compared with the results of the present empirical formula and the analytical formula of
Wiegel (1961), as well as physical model data (Madsen 1991; Li and Yan 2005; Fuhrman and
Madsen 2008).

2 Existing analytical and empirical formulas for transmission


coefficients and present formulas
The transmission coefficient is directly related to the type of wave-permeable structures.
The two types primarily examined in this investigation were the emergent porous rubble
mound type (Fig. 1(a)) and pile type (Fig. 1(b)).

Fig. 1 Permeable breakwaters

458 Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467
For a rubble mound breakwater, based on extensive experimental data, van der Meer and
Daemen (1994) suggested the following expression using the medium diameter of breakwater
body stones to describe the transmission coefficient:
R
Kt = a c + b (1)
Dn50
where K t is the transmission coefficient, with K t = H k H s ; H k is the height of the
transmitted wave behind the permeable breakwater; H s is the significant wave height of the
incident wave; Rc is the crest freeboard; Dn50 is the medium diameter; a and b are parameters,
with a = 0.031H s Dn50 − 0.024 , and b = −5.42Sop + 0.0323H s Dn50 − 0.017 ( B Dn50 )
1.84
+ 0.51 ;
Sop is the wave steepness, with Sop = 2πH s ( gT ) ;
p
2
Tp is the peak period of the incident
wave; g is the gravitational acceleration; and B is the upper structure width.
d’Angremond et al. (1996) related the transmission coefficient to Rc H s , B H s , and
breaker parameter ζ op :
−0.31
R  B 
K t = −0.4 c + 0.64   1 − exp ( −0.5ζ op )  (2)
Hs H  
 s
where the breaker parameter or the Iribarren number ζ op = hx ( Sop ) , and hx is the seaward
12

slope of the structure.


For the rubble mound breakwater, laboratory test results are available from various
studies (Kramer et al. 2005; Calabrese et al. 2008; Peng et al. 2009; Hur et al. 2012; Melito
and Melby 2002; Laju et al. 2011). They were compared with the two formulas above in terms
of function design of the crest freeboard Rc and upper structure width B. As a result, Eq. (1)
can be applied to the rubble mound breakwater; Eq. (2) is not limited to the rubble mound
breakwater, but a large deviation may occur when it is applied to other types of permeable
breakwaters, especially those with a large transmission coefficient.
Based on the small amplitude wave theory, the analytical solution of linear waves across
a thin vertical pile interface was suggested by Wiegel (1961). The expression based on the
linear theory neglects the nonlinear damping effect, and thus the equation provides a larger
transmission coefficient:
1/ 2
 2k ( d − Rt ) + sinh 2k ( d − Rt ) 
Kt = η   (3)
 2kd + sinh 2kd 
where the adjustment coefficient η is less than 1, and, usually, η = 0.9 according to a
simple correlation analysis considering energy dissipation; k is the wave number; d is the still
water depth; and Rt is the submerged depth of a splashing board of pile-type breakwaters.
The present investigation was performed based on available data and above analytical
and empirical formulas. For the rubble mound breakwater, the crest freeboard Rc plays an
important role in evaluating transmission effects, which is crucial in determining wave
transmission, reflection, and overtopping, and has a linear relation with K t in Eq. (2).

Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467 459
Considering this factor, a better way to correlate Rc with K t using a shape function is
R
1 − α1 c
Hs  B 
K t = β1
Rc
exp  −0.18  1 − exp ( −0.5ζ op )  (4)
 Hs 
1 + α1
Hs
where α1 (here α1 Rc H s ≤ 1.0 ) and β1 are two coefficients for the rubble mound breakwater.
Since Rc H s is usually close to 1 in practice, with reference to the Code of Design and
Construction of Breakwaters (MT PRC 2011), we determined that α1 = 1.0 and β1 = 0.90 by
fitting the laboratory data from van der Meer et al. (2005) and conducting a simple correlation
analysis, as shown in Fig. (2).

Fig. 2 Determination of α1 and β1 for emergent porous rubble mound breakwater by fitting data extracted
from van der Meer et al. (2005)
For the pile-type permeable breakwater, the submerged depth of the splashing board Rt
plays an important role in defining the transmission coefficient K t . Replacing α1 Rc H s in
Eq. (4) with α 2 Rt H s , the following empirical formula is introduced:
R
1 − α2 t
Hs  B 
K t = β2 exp  −0.18  (5)
Rt  Hs 
1 + α2
Hs
where α 2 and β 2 are two coefficients for the pile-type breakwater, which were determined
to be 0.23 and 2.3, respectively, through analysis of the experimental data (Li and Yan 2005)
provided in the next section.
The submerged breakwater (Fig. 1(d)) is more similar to the pile-type breakwater than to
the rubble mound breakwater in wave transmission effects, since both the pile type and the
submerged type have a continuous water body. As shown in Fig. 3, α 2 and β 2 were
determined to be 0.23 and 0.50, respectively, in Eq. (5) for the submerged breakwater,
based on fitting of the laboratory data from van der Meer et al. (2005) and a simple
correlation analysis.

460 Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467
Fig. 3 Determination of α 2 and β 2 for submerged breakwater by fitting data extracted from
van der Meer et al. (2005)

3 Comparison of present and existing empirical formulas of


transmission coefficients
Based on 2 337 LCS data sets from van der Meer et al. (2005) and Zanuttigh et al.
(2008), Eq. (4) was compared with Eq. (2). As shown in Fig. 4(a), Eq. (4) has a wider range of
application, especially for B H s >10 . Fig. 4(b) indicates that the transmission coefficient K t
decreases exponentially with the increment of Rc H s according to Eq. (4), which is more
consistent with the measured data than the linear relation in Eq. (2).

Fig. 4 Comparison of calculated results of Eq. (2) and Eq. (4) for rubble mound breakwater

As described by Fig. 5, for a pile-type breakwater, the calculated results of Eq. (3) show
larger values of the transmission coefficient K t than the results of Eq. (5) and laboratory data
from Li and Yan (2005), because Eq. (3) does not consider energy dissipation by LCS. We
determined that α 2 = 0.23 and β 2 = 2.3 in Eq. (5) based on a simple correlation analysis.
Fig. 5 shows that the calculated results of Eq. (5) are closer to the laboratory data from Li and
Yan (2005). For larger values of relative submerged depth Rt H s (i.e., smaller values of
Rc H s ), overtopping less than 0.2 H s may occur in the laboratory data, while, for smaller
values of Rt H s (i.e., larger values of Rc H s ), experimental data from Li and Yan (2005)

Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467 461
show no overtopping effects. It is suggested that the overtopping effects for larger values of
Rt H s may be considered by slightly increasing the value of coefficient β 2 in Eq. (5). In
conclusion, Eq. (5) shows an acceptable accuracy in comparison with experimental data and
can be used for estimating the transmission coefficient for pile-type breakwaters.

Fig. 5 Comparison of experimental data (Li and Yan 2005) and calculated results of
Eq. (3) and Eq. (5) for pile-type breakwater

4 Numerical flume tests based on MBE model and verification of


present empirical formula for pile-type breakwater
4.1 MBE model

In the numerical flume, the friction term of the porosity expression was added to
governing equations to modify the classic Boussinesq-type wave equations (Madsen 1983).
The nonlinear friction term inside the permeable structure (or the porous absorber) in shallow
water can be linearized by using the following approximation:
ω
U (α + β U ) ≅ f
U (6)
n
where U is the velocity vector; α and β are the friction coefficients of permeable structures
accounting for the laminar and turbulent friction losses, respectively; f is the friction factor
independent of time t and space in x and y dimensions; ω is the incident wave angular
frequency; and n is the porosity coefficient of structures. α and β are related to the
porosity coefficient n from Engelund (1953):
(1 − n ) υ
3

α = α0 (7)
n2 2
Dn50
(1 − n ) 1
β = β0 (8)
n3 Dn50
where υ is the kinetic viscosity, and α 0 and β 0 are constants.
The expression on the right-hand side of Eq. (6) (Madsen 1983) was derived under
one-dimensional shallow-water long-wave conditions by neglecting higher order nonlinear
effects. Considering that this assumption causes uncertainty in understanding short wave

462 Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467
transmission, the items on the left-hand side of Eq. (6) are directly applied to MBE by setting
β as a constant in numerical flume tests and α at a value less than 10. Eqs. (7) and (8)
demonstrate the porosity effects on the hydrodynamic characteristic of permeable breakwaters
by assuming a nominal Dn50 value.
Since the MBE model is based on the depth-averaged two-dimensional water flow, the
corresponding mathematical formulation of permeable structures is similar to that used to
solve the problem of waves passing through permeable screens, where the transmitted and
reflected waves occur on both sides of the screen. In the numerical flume, the transmission
coefficient can be determined by the friction factor α and the number of friction layers (Li
and Yan 2005; Zhang and Li 2008). Different types of permeable structures, as shown in Fig. 1,
were tested with the MBE model in this study.

4.2 Comparison of numerical flume results with physical model data

The wave attenuation effects were tested for four specific types of permeable breakwaters
using the MBE model. The instant wave elevation distribution and wave height distribution for
waves moving through different types of permeable breakwaters are illustrated in Fig. 6 and
Fig. 7. A typical standing wave and a transmitted propagating wave form respectively in front
of and behind the permeable structures for all types of breakwaters except for the porous
breakwater before a rigid wall. The numerical flume results using the MBE model and
physical model data for different types of permeable breakwaters are listed in Table 1,
where K r = H r H s , with H r indicating the reflected wave height in front of the breakwater.
It can be concluded that the numerical flume results agree with the physical model data.

Fig. 6 Wave envelopes in numerical flume

Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467 463
Fig. 7 Wave height distributions in numerical flume
Table 1 Comparison of numerical flume results using MBE model with physical model data for
different types of permeable breakwaters
Kr Kt
Breakwater α Data
n Physical Physical
type MBE model MBE model source
model model
P 1.50 0.75 0.35 0.34 0.30 0.32 Chen et al. (2008)
E 0.30 0.86 0.12 − 0.22 0.22 van der Meer et al. (2005)
R 9.53 0.63 0.55 0.56 0 0 Madsen (1983)
S 0 0 0.03 0.02 0.54 0.55 Luth et al. (1994)
Note: P, E, R, and S mean the pile-type breakwater, emergent porous rubble mound breakwater, porous breakwater before a rigid
wall, and submerged breakwater, respectively.

4.3 Comparisons of numerical flume results with physical model data and
results of analytical and empirical formulas for pile-type breakwater

Numerical flume tests on the pile-type breakwater were performed using the MBE model.
The data sets from P01 to P12 for the numerical flume are listed in Table 2, where B H s is
determined to be 5 in accordance with the physical model tests (Chen et al. 2008).
Table 2 Data sets for numerical flume tests
Data set Rt H s α n Data set Rt H s α n
P01 –0.20 0.1 0.90 P07 1.82 1.4 0.77
P02 0.37 0.3 0.86 P08 2.05 1.8 0.75
P03 0.55 0.4 0.83 P09 2.50 2.8 0.69
P04 0.85 0.5 0.82 P10 2.70 3.5 0.65
P05 1.00 0.6 0.81 P11 2.90 4.2 0.60
P06 1.20 0.8 0.80 P12 2.95 4.5 0.57

The numerical flume results were compared with the calculated results of Eq. (5) and
Eq. (3) as well as the physical model data from Chen et al. (2008), as shown in Fig. 8. A
decaying trend of K t is demonstrated with the increment of Rt H s , and the empirical curve
of Eq. (5) agrees with the numerical flume results and the measured data. In general, the
overtopping effects of the pile-type breakwater will lead to an increase of K t . This is outside
of the scope of the present investigation since these effects only occur in physical models. By
fitting the data of α and Rt H s in Table 2, as shown in Fig. 9, a new empirical relation

464 Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467
relating α to Rt H s was obtained as follows:
 Rt 
α = 0.23exp   (9)
 Hs 
Comparisons of the transmission coefficient K t obtained by the MBE model with those
obtained by the formulas (Eq. (3) and Eq. (5)) and the physical model against the friction
coefficient α and the porosity coefficient n, with the friction coefficient α obtained by
Eq. (9), are shown in Fig. 10. The agreement between the results of the MBE model, the
physical model, and Eq. (5) demonstrates that Eq. (9) has a strong fitting effect, and the
numerical flume has a wide range of applicability. The results also show that Eq. (5) is
superior to Eq. (3) in estimating the transmission coefficient for pile-type breakwaters.

Fig. 8 Comparisons of K t against Rt H s Fig. 9 α versus Rt H s

Fig. 10 Comparisons of K t against α and n 

4.4 Discussion
In coastal engineering practice, Rt H s may exceed 2 for many reasons, such as
reduction of the transmission coefficient K t by increasing reflection and regulating tidal
levels. In addition, damping effects of B H s play an important role in breaking waves, which
means that a close relationship exists between K t and B H s : as Rt H s → 0 in Eq. (9),
α → 0.23, and, accordingly, K t → 2.3exp ( −0.18 B H s ) in Eq. (5). Considering that
viscosity and nonlinearity are of great significance in the determination of B H s , especially

Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467 465
for three-dimensional shallow water, further verification and improvement regarding the
nonlinearity should be focused on the relationship between B H s and the coefficient β in
Eq. (6) in the future. Although numerical simulation tests are becoming more valuable with the
development of computer capability and calculation speed, the MBE model can only be
successfully applied to two-dimensional simulations, and three-dimensional simulations of
wave phenomena are still difficult. Thus, physical models still play an important role in
understanding basic mechanics of wave phenomena.

5 Conclusions
Based on the analysis of existing analytical and empirical formulas for the transmission
coefficient of permeable breakwaters, new empirical formulas were developed, with their
coefficients obtained by fitting laboratory data. These formulas were verified through
comparison with numerical flume tests and physical model tests. It can be concluded that the
present formula for rubble mound-type breakwaters has a wider application range than the
existing empirical formulas, and that the results of the present formula for pile-type
breakwaters are more consistent with both the numerical flume results and the physical model
data than those of the existing analytical formula. Moreover, an empirical formula relating the
friction coefficient α to the relative submerged depth Rt H s was also derived and verified,
in which α is a vital factor in wave transmission simulation with the MBE model.

References
Calabrese, M., Buccino, M., and Pasanisi, F. 2008. Wave breaking macrofeatures on a submerged rubble
mound breakwater. Journal of Hydro-environment Research, 1(3-4), 216-225. [doi:10.1016/10.1016/
j.jher.2007. 11.003]
Chen, D. C., Wang, Y. G., Li, X., Yang, Y., and Cai, H. 2008. Experimental Study of Wave Permeable
Breakwater in Sansha Central Fishing Port. Nanjing: Hohai University. (In Chinese)
Christou, M., Swan, C., and Gudmestad, O. T. 2008. The interaction of surface water waves with submerged
breakwaters. Coastal Engineering, 55(12), 945-958. [doi:10.1016/j.coastaleng.2008.02.014]
d’Angremond, K., van der Meer, J. W., and de Jong, R. J. 1996. Wave transmission at low crested structures.
Proceedings of the 25th International Conference on Coastal Engineering, 2418-2427. Florida: ASCE.
Engelund, F. 1953. On the laminar and turbulent flows of ground water through homogeneous sand.
Transactions of Danish Academy Technical Science, Vol. 3.
Fuhrman, D. R., and Madsen, P. A. 2008. Simulation of nonlinear wave runup with a high-order Boussinesq
model. Coastal Engineering, 55(2), 139-154. [doi:10.1016/j.coastaleng.2007.09.006]
Hur, D. S., Lee, W. D., and Cho, W. C. 2012. Characteristics of wave run-up height on a sandy beach behind
dual-submerged breakwaters. Ocean Engineering, 45, 38-55. [doi:10.1016/j.oceaneng.2012.01.030]
Kramer, M., Zanuttigh, B., van der Meer, J. W., Vidal, C., and Gironella, F. X. 2005. Laboratory Experiments
on low-crested breakwaters. Coastal Engineering, 52(10-11), 867-885. [doi:10.1016/j.coastaleng.
2005.09.002]
Laju, K., Sundar, V., and Sundaravadivelu, R. 2011. Hydrodynamic characteristics of pile supported skirt
breakwater models. Applied Ocean Research, 33(1), 12-22. [doi:10.1016/j.apor.2010.12.004]
Li, C. L., and Xie, Y. Y. 2008. Random wave motions around different submerged dikes. Coastal Engineering,
27(1), 1-9. (in Chinese)
Li, X., and Yan, Y. X. 2005. Numerical simulations of nonlinear wave transformation around wave-permeable

466 Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467
structure. Journal of Hydrodynamics, Ser. B, 17(17), 699-703.
Luth, H. R., Klopman, Q., and Kitou, N. 1994. Kinematics of Waves Breaking Partially on an Offshore Bar.
Delft: Delft Hydraulics.
Madsen, P. A. 1983. Wave reflection from a vertical permeable wave absorber. Coastal Engineering, 7(4),
381-396. [doi:10.1016/0378-3839(83)90005-4]
Madsen, P. A. 1991. A new form of the Boussinesq equations with improved linear dispersion characteristics.
Coastal Engineering, 15(4), 371-388. [doi:10.1016/0378-3839(92)90019-Q]
Melito, I., and Melby, J. A. 2002. Wave runup, transmission, and reflection for structures armored with
CORE-LOC. Coastal Engineering, 45(1), 33-52. [doi:10.1016/S0378-3839(01)00044-8]
Ministry of Transport of People’s Republic of China (MT PRC). 2011. Code of Design and Construction of
Breakwaters (JTS 154-1-2011). Beijing: China Communications Press. (in Chinese)
Peng, Z., Zou, Q., Reeve, D. E., and Wang, P. X. 2009. Parameterisation and transformation of wave
asymmetries over a low-crested breakwater. Coastal Engineering, 56(11-12), 1123-1132. [doi:10.1016/
j.coastaleng.2009.08.005]
van der Meer, J. W., Briganti, R., Zanuttigh, B., and Wang, B. X. 2005. Wave transmission and reflection at
low-crested structures: Design formulae, oblique wave attack and spectral change. Coastal Engineering
52(10-11), 915-929. [doi:10.1016/j.coastaleng.2005.09.005]
van der Meer, J. W., and Daemen, I. F. R. 1994. Stability and wave transmission at low crested rubble mound
structures. Journal of Waterway, Port Coastal and Ocean Engineering, 120(1), 1-9. [doi:10.1061/(ASCE)
0733-950X(1994)120:1(1)]
Wiegel, R. L. 1961. Closely spaced piles as a breakwater. Dock and Harbor Authority, 42(491), 150.
Zanuttigh, B., Martinelli, L., and Lamberti, A. 2008. Wave overtopping and piling-up at permeable low crested
structures. Coastal Engineering, 55(6), 484-498. [doi:10.1016/j.coastaleng.2008.01.004]
Zhang, S. X., and Li, X. 2008. Numerical simulation of nonlinear wave propagating in flume. Proceedings of
Fourth International Conference on Natural Computation, 649-653. Perth: IEEE. [doi:10.1109/ICNC.
2008.854]
(Edited by Ye SHI)

Su-xiang ZHANG et al. Water Science and Engineering, Oct. 2014, Vol. 7, No. 4, 457-467 467

Você também pode gostar