This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
and £. A n iterative solution of the equation for ip is applied to obtain <j> as a function of ip and the subsequent use of 4>{(p) in the equation for ii!s yields an expression for ws in terms of ( a n d (p. This expression for ws is now used in the Fourier transform of the free-surface equations (3). Multiplying t h e equation for C, by y/g/(2uj(k)) a n d multiplying t h e equation for
, Li and Anastasiou (1992) suggest that as few as 2 or 3 grid points per wavelength will suffice. However, as noted by Li and Anastasiou (1992) and by Radder (1992), the presence of rapidly varying topography or of reflections in various directions will necessitate much finer resolution (say 10 points per wavelength) and the solution obtained by using the log of the potential may lead to excessive smoothing. For mildly varying bathymetry with low reflections, it may in any case be more efficient to solve instead of the "parabolic approximation" of Eq. (1) which is intended for such applications. Li and Anastasiou (1992) have used the multigrid method to minimize storage problems. When higher resolution is required, though, this method does not offer any significant advantage over the conjugate gradient schemes described by Li (1994a) and Panchang et al. (1991) since at least one grid with the desired resolution must be constructed. Further, the multigrid method is best suited to rectangular finite-difference discretizations. Another method, proposed by Li (1994b), involves solving the following parabolic equation, a^ xfD)>P = 0,
= V • (CCgV
(29)
where a is a constant. Equation (29) is an approximation of the time-dependent hyperbolic wave equation associated with Eq. (1). It is solved by marching forward in time until steady state is reached. Equation (29) is similar to the heat equation and standard techniques (e.g., the ADI method) for solving such equations are used by Li (1994b). It must be noted, though, that the elimination
142
V. Panchang & Z.
Demirbilek
of equations. For conjugate gradient solvers, 90% of the CPU time is spent on matrix-vector products and inner product kernels. Therefore, OpenMP (OARB 1997) may be used to parallelize the kernels. Two-level parallelization schemes can use OpenMP to accelerate the solution for each component and MPI to simultaneously obtain solutions to multiple incident wave components. More details regarding parallelization schemes for harbor wave models may be found in Bova et al. (2000) who report a reduction in run times by a factor of 250-580 compared with serial codes for an application in Ponce de Leon Inlet (Florida). A problem with nearly 300 input spectral components was solved on a 25 square km domain containing 235,000 nodes in 72 hours. An example from Bova et al. (2000), shown in Fig. 6, suggests that the simulation produces, qualitatively, a sea-surface that looks realistic. Model results for this site are discussed in greater detail by Zhao et al. (2000). 4. Incorporation of Additional Mechanisms As noted earlier, Eq. (1) incorporates the effects of refraction, diffraction and reflection induced by any nonhomogeneity in the model domain. Equation (5) is an extension of Eq. (1) that includes, in addition, the effects of friction and wave breaking. Similar extensions are possible to include the effects of wavecurrent interaction, wave-wave interaction and of steep slopes. The modeling of these mechanisms in the context of the elliptic equation, Eq. (1), is described in this section. 4.1.
Dissipation
In Eq. (5), W represents the combined effects of friction and breaking which may be separated as follows, W = w/Cg+j,
(30)
where w is the friction coefficient defined by Dalrymple et al. (1984) and 7 is a breaking factor. These coefficients are empirical and parametrizations for these have been described by Dalrymple et al. (1984), Tsay et al. (1989) and Chen (1986) for friction, and by Battjes and Janssen (1978), Dally et al. (1985), Massel (1992), Chawla et al. (1998) and Isobe (1999) for breaking. Some of these parameterizations have been extensively validated against field data (e.g., Larson, 1995; Kamphuis, 1994). We do not repeat the parameterizations here; rather, we note that they are all dependent on the wave amplitude.
Simulation
of Waves in Harbors
143
Published studies demonstrating the effects of friction in harbor models (e.g., Chen, 1986; Tsay et al, 1989; Demirbilek and Panchang, 1998; Kostense et al, 1986) have estimated w on the basis of the incident wave amplitude. It is then easy to pre-specify w while solving Eq. (5). These studies appear to show that friction can change the magnitude of resonant peaks in harbor models quite substantially; at other frequencies, the effect seems to be minimal. Jeong et al. (1996) and Moffatt & Nichol Engineers (1999) have attempted to utilize W to include harbor entrance losses, however, no details regarding the modeling technique are presented. In general, however, since both w and 7 are functions of the wave amplitude which is unknown a priori inside the domain, their inclusion makes the problem nonlinear and requires iteration. For the first iteration, W is set equal to 0 and Eq. (1) is solved (e.g., nonbreaking solutions are obtained). The resulting wave heights are used to estimate W via the parameterizations for w and 7 and Eq. (5) is solved. The process is repeated until convergence is obtained. Since dissipation (especially breaking) occurs outside the computational domain also, open boundary conditions like Eqs. (13), (15) and (16) may not be appropriate. Inclusion of breaking inside the domain and its exclusion in the exterior descriptions create artificial discontinuities along the open boundary, especially in shallow areas, and consequently, spurious effects would propagate into the model domain. In this event, Eq. (20) is a more appropriate description of the exterior and may be used to develop the necessary boundary conditions. A digitized bathymetry file is used to obtain the depths d(x) along transect 1. These depths are interpolated onto uniformly spaced nodes and the wave properties C, Cg and k are calculated. Equation (20) may be easily solved by finite-differences using boundary conditions, Eqs. (24) and (25). Again, iterations are required because W is not known initially. When the solutions converge, the procedure is repeated for transect 2. These converged solutions of Eq. (20) along transects 1 and 2 include, albeit in a one-dimensional sense, the effects of dissipation and hence constitute more appropriate forcing functions than Eqs. (13), (15) and (16) do. 0o along the semicircular open boundary is obtained via Eq. (26). Performing nonlinear iterations within the model domain as W varies from iteration to iteration can be time-intensive. We have explored the possibility of combining the iterative conjugate gradient methods for the linear system with the iterations required for the nonlinear modeling, i.e., the conjugate gradient iterates obtained while solving Eq. (28) were perturbed by upgrading W
144
V. Panchang & Z.
Demirbilek
periodically. Unfortunately, this maneuvering destroys the robust convergence properties of the conjugate gradient solvers for Eq. (28). At present, each linear system, for a specified W, must be completely solved until convergence is obtained and the whole procedure repeated with a new W. More effective methods to accelerate the solution need to be developed. Zhao et al. (2000) developed a finite-element model using Eqs. (6), (20) and (27) to formulate the boundary conditions and applied it to several tests involving breaking. These tests involved a sloping beach, a bar-trough bottom configuration, shore-connected and shore-parallel breakwaters on a sloping beach, and two field cases in the North Sea and Ponce de Leon Inlet (Florida). Five breaking formulations, given by Battjes and Janssen (1978), Dally et al. (1985), Massel (1992), Chawla et al. (1998) and Isobe (1999), were examined. They found that the Isobe (1999) criterion was difficult to use within the context of the elliptic model and that the absence of a lower breaking limit generally contributed to excessive dissipation (compared with data) in the Chawla et al. (1998) and Massel (1992) formulations. In general, the formulations of Battjes and Janssen (1978) and Dally et al. (1985) were found to be the most robust from the point of view of incorporation into an elliptic model based on Eq. (5) and to provide excellent results compared to data. For simulations involving several spectral components, Zhao et al. (2000) examined two approaches. In the first approach, complete simulations were made one at a time for all monochromatic components where the amplitude of each component was used in the relevant breaking formula. The results were subsequently assembled using linear superposition. They found that this approach led to some overestimation compared with data; this was attributed to the individual component amplitudes being too small to induce breaking in the model. In view of this overestimation, a second approach was considered where the breaking factor was calculated on the basis of the significant wave height instead of the component wave height. This approach eliminates the independence of individual component simulations, thereby changing the overall model numerics. With the second approach, one round of nonlinear iterations for all components must be performed, the significant wave height is calculated at each grid point and this larger wave height is used to estimate the breaking factor (Chawla et al, 1998). This approach led to the initiation of breaking occurring further offshore in the case of their simulations of wave transformation around Ponce de Leon Inlet. An example of the model simulations near the US Army Field Research Facility at Duck, North Carolina, is shown in Fig. 7 for an input wave condition given by a significant wave height of 2.3 meters.
Simulation
of Waves in Harbors
145
Fig. 7. Modeled wave amplitudes (m) at F R F Duck. Top, no breaking; Bottom, breaking based on significant wave height.
A complex pattern of waves is created in the middle of the domain due to the complicated bathymetry. Clearly, breaking plays an important role (the dots aligned in the shore-perpendicular direction in the middle of these figures represent circular piles which were assigned full reflection). These simulations were performed with 208 spectral components for a storm in 1996. The
146
V. Panchang & Z.
Demirbilek
simulations took 265 hours CPU time using 41 processors and ran nonstop for about 3 days on the US Army Corps of Engineers super-computer. Comparison to field data and other details are provided elsewhere.
4.2. Wave-current
interaction
Many coastal regions experience high background currents. Wave propagation is influenced by these currents (e.g., waves opposing the currents become larger and vice-versa). Based on the derivation by Kirby (1984), one may incorporate currents in Eq. (5) as follows, V • (CCgV
+ iCgarW
+ a2 - a2 + iaV • U (31)
where \J(x, y) = current vector (provided by a flow model), oy = a — k • U = relative frequency, {a2 = gk tanh (kd); C — ar/k). Several sophisticated hydrodynamic models are available nowadays for obtaining the desired flowfield information TJ(x,y). When hydrodynamic models provide three-dimensional flowfields, the vertical dependence may be removed for use in Eq. (31) via the "equivalent uniform current" defined by Hedges and Lee (1992); this quantity is obtained by vertically averaging the current over a depth eL over the water column where eL = (1/k) tanh (fed). The generalized mild-slope wave equation, Eq. (31), is still elliptic and may be solved by the techniques noted previously. For a prespecified U(x, y), a linear system of equations like Eq. (28) results. However, to compute the Doppler shift in the wave frequency (oy = a — k • U), the wave vector k is needed. While the magnitude of k is known a priori from the dispersion relation, Eq. (2), its direction is not. This problem may be resolved by first solving Eq. (31) without the effects of wave-current interaction, obtaining an estimate of the local wave direction, computing the relative frequency ar = a — k • U, solving Eq. (31) again, revising the wave direction, and repeating until the model runs converge. Such an approach has been taken by Kostense et al. (1988); Li and Anastasiou (1992), however, prefer not to calculate the direction of k in view of the computational burden. While their results for one test-case (pertaining to waves approaching a rip current on a sloping beach) are reasonable, iterations are indeed necessary for complex flowfields. Further difficulties arise in the specification of open boundary conditions. Published studies using the elliptic model of Eq. (31) have assumed that
Simulation
of Waves in Harbors
147
currents are absent on this boundary. Research is needed to develop open boundary conditions that include the effects of currents (an extension of Eq. (20) is possible for the case of currents varying only in the cross-shore direction outside the computational domain). 4.3. Wave-wave
interaction
By including most of the nonlinear terms in the vertical integration of the three-dimensional Laplace equation, Kaihatu and Kirby (1995) obtained an extension of Eq. (1) that incorporates wave-wave interactions. Expressing the potential in terms of harmonics as: N
^2fn(kn,d,z)(j)n(kn,ujn,x,y,t).
(32)
71 = 1
where, _ coshkn(d + z) In — cosh knd
(33)
and performing an integration over the vertical modifies Eq. (5) as follows, Vfc • [(CCg)nV
kl(CCg)n
"ra-l
J2 2(TnV
n— 1 "i" Gl<j>l^
$11—1 "i" &n— l^n
—1^
n
2
N-n
2cr„V!>,* • Vcj>n+i + an+i4>n+iV24>*
^
-
(Ti(j>*V24>n+i
L i=i
2
(°"/ ~
(34)
complex conjugate value of < in which q Similar equations were also derived by Tang and Ouellet (1997) who have further demonstrated that this type of extension provides the governing equation, Eq. (1), with the same level of nonlinearity as that in Boussinesq wave models. This is particularly noteworthy since models based on Eq. (1) or
148
V. Panchang & Z.
Demirbilek
/ / / / f /
/
0.45 ( 0.35 ( 0.25 ( 0.15
\ \ V \ \ \. \. A = 0.98 cm T = 3s
0.15 m
±_
t
0.46 m
(a)
Model
• Data
Model
• Data
0.02 0.015
Second Harmonic
•S 0.01 0.005
•
0 20
25
0.02 — Model 0.015
-
• Data
Third Harmonic
§ 0.01 0.005 0
=ȣ*** 10
15
20
25
x(m)
(b) Fig. 8. (a) Bathymetry (m) of Whalin (1971) used for wave-wave interaction study, (b) wave height comparison for wave-wave interaction.
Simulation
of Waves in Harbors
149
Eq. (5) simultaneously offer the computational stability and the advantages of finite-element gridding in harbors and complex coastal areas (which Boussinesq models sometimes lack). From the perspective of the solution technique, the coupling of harmonics represented by the right hand-side of Eq. (34), if prespecified, leads to a linear system of equations like Eq. (28). However, for a given component <j)n, the other components constituting the right hand side are not known a priori. Again, an iterative technique must be used where the values from the previous round can be used to calculate the right hand side. Figure 8 shows a finite element model simulation (based on Eq. (34)) of wave propagation and interaction over the "tilted cylinder" bathymetry of Whalin (1971). The results match the laboratory data of Whalin (1971) very well and shows that higher harmonics can build up from zero to a magnitude similar to the linear solution and can hence contribute much to the overall solution (hitherto the coupling represented by the right hand side was included only in simple (parabolic approximation) models; e.g., Tang and Ouellet, 1997 and Kaihatu and Kirby, 1995). 4.4. Combined
nonlinear
mechanisms
Equation (34) contains the effects of two of the additional mechanisms described so far, i.e., wave-wave interactions and dissipation. By rederiving Eq. (34) on a moving frame of reference, the equation may be extended to simultaneously include the effects of wave-current interaction also (Kaihatu and Kirby, 1995). As demonstrated above, the modeling procedures for each of these mechanisms are individually nonlinear and require numerical iterations. However, combining all the nonlinear effects in numerical simulations has as yet been unexplored. An efficient model must juxtapose iterations and also assure convergence. Further, appropriate tests for the enhanced model are not readily available (especially for the combination of wave-wave and wave-current interactions). For systematic model verification, data isolating and combining these mechanisms are needed. In that regard, the numerical advancements appear to be preceding data availability. 4.5. Steep-slope
effects
Unlike the inclusion of the nonlinear mechanisms described above, overcoming the "mild slope" requirement discussed in Sec. 1 is relatively easy. Massel (1993), Porter and Staziker (1995), Chamberlain and Porter (1995), and
150
V. Panchang & Z.
Demirbilek
Chandrasekera and Cheung (1997) developed extensions of Eq. (1) to include steep-slope effects. Their extensions may be described by the following equation, V • (CCgV(f>) + {k2CCg + dx (V/i) 2 + d2V2h)(j> = 0,
(35)
where d\ and d2 are functions of local depths. Reference may be made to these publications for the various definitions of d\ and d2; in general, though, differences in the proposed definitions of these functions impact model results to a very small extent. The steep-slope terms are fairly straightforward to include in the model because they are linear. Further, they have the advantage of being "automatic", i.e., they have little contribution for mild slopes, do not change the solution technique and the additional computational demand is negligible. However, steep slopes lead to breaking and model performance in the vicinity of steep slopes will involve iterations (an analytical model has been developed by Massel and Gourlay (2000) to include breaking and steep-slope effects near coral reefs). 5. Application to Harbors So far, we have described various developments made in recent years to construct more reliable models based on Eq. (1) for use in domains with arbitrary shape and bathymetry. In this section, we describe application of one such model to the practical problem of simulating harbor resonance in the Los Angeles/Long Beach Harbor complex (Fig. 1) and in Barber's Point Harbor (Hawaii). Both harbors are undergoing considerable renovation to accommodate increased shipping. A finite-element model called CGWAVE was developed to solve Eq. (5) using Eqs. (6), (18), (20) and (27) to formulate the boundary conditions. The Surface Water Modeling System (Zundell et al., 1998) was used for grid generation. 5.1. Simulations in the Los Beach Harbor complex
Angeles/Long
The Los Angeles/Long Beach Harbor complex (Fig. 1) is one of the largest harbors in the world; therefore, the model domain is quite large covering an area of approximately 120 square km. Bathymetric input was obtained by digitizing NOAA chart number 18749. For numerical modeling, a grid containing 285,205 triangular finite elements was developed. It was based on a resolution
Simulation
of Waves in Harbors
151
of 10 points per wavelength for a 30-second wave. The two one-dimensional transects in the exterior were extended in the offshore direction to a distance of 9.2 km beyond which the depth was assumed to be constant. At this location, the input wave was specified. For initial quality control simulations, the coastal reflectivity was initially set equal to zero (i.e., fully absorbing) since this case is easier to examine qualitatively than the case when a large number of reflections are present. Figure 9 shows the phase diagram for a 50 second wave. The results appear to be quite satisfactory. A reduction in the wavelength in the onshore direction is evident. No spurious boundary effects are seen. Penetration through the breakwater gaps is precisely as one would expect. Bending of the crests as they approach from onshore also indicates a correct reproduction of refractive effects. For further simulations, the coastal boundary was assumed to be fully reflecting (both inside the model and for the one-dimensional transects). Also, the geometry of the offshore breakwaters was changed. These breakwaters are known to be permeable to waves (e.g., Chiang, 1987) and it is hence not appropriate to consider them as closed boundaries. Permeable structures cannot be easily handled within the context of an elliptic boundary value problem.
Fig. 9. Modeled phase diagram for Los Angeles/Long Beach Harbor complex; 50 second obliquely incident wave.
152
V. Panchang & Z.
Demirbilek
One approach may be to treat the breakwater as a water area and ascribe an appropriate dissipation factor in that region. We used an alternative approach whereby the breakwater was divided into several segments so that energy could propagate through gaps in the breakwater. 50% of the breakwater length was opened up by means of numerous gaps interspersed among several solid segments. Seabergh and Thomas (1995) conducted hydraulic model simulations for this complex at the US Army Waterway Experiment Station in Vicksburg, Mississippi. They collected data at several gages, shown in Fig. 1, for various harbor plans. The bathymetric data used for numerical modeling obtained from the more recent NOAA chart was a reasonable approximation of the harbor geometry described as "Stage II" by Seabergh and Thomas (1995). However, neither the bathymetry data nor the boundary geometries used in the two studies were identical. Seabergh and Thomas (1995) performed their hydraulic model experiments for a large number of input frequency components varying from 30 seconds to 512 seconds. At each gage location, the amplification factor was measured for several frequencies and a resonance curve was developed. These curves were found to be extremely noisy, i.e., the response varied quite rapidly with frequency at the gages (see example in Fig. 10). For convenience of analysis, therefore, they partitioned the data into three groups: short period waves (30 s Gage 56 Resonance Curve 16 14
J! 12 10 8
a.
e <
6 4 2 0 10
100 Wave Period (s) Fig. 10. Resonance curve at Gage 56.
1000
Simulation
of Waves in Harbors
153
to 42 s), medium period waves (42 s to 205 s) and long period waves (205 s to 512 s). For each gage, the amplification factors within each group were averaged over the respective frequencies. Numerical simulations were performed for three incident angles, for normal incidence and for 30° on either side of it, to account for the effects of the wave maker. The results of the three directional inputs were averaged for each frequency. The exact location of each gage was not known, so results in the general vicinity of the gage as determined from Fig. 1 were extracted and averaged over the frequency bands stated earlier. In all, simulations were made for 10, 30 and 17 frequency components in the three bands. These components are irregularly spaced and correspond approximately to the discrete frequency components used by Seabergh and Thomas (1995) in their hydraulic model simulations. An example of the modeled resonance curve is shown in Fig. 10. At T = 45 s, the lab data show a remarkably high amplification that the model underpredicts; conversely, for T between 300 s and 400 s, the model value is greater than the hydraulic model data. The overall results for all gages, using the averaging described above, are compared against the hydraulic model data in Fig. 11. In general, the numerical simulation predicts the response at the gages as well as the hydraulic model data. The agreement is quite good for the short and medium period waves. Greater discrepancy is seen for the long waves which also exhibit greater gage-to-gage variability. For the long waves, there seems to be systematic overprediction near certain gages. These discrepancies could be attributed to several factors. First, the two bathymetry sets are not identical and the high variability implies that small differences in the geometry can result in large differences in the response. The location of the input wave is also different in the hydraulic and numerical models. Further, the exterior sea is bounded in the hydraulic model, thus, possibly preventing radiation out to the open sea. Finally, reflection coefficients and the degree of permeability of the breakwaters are not sufficiently well-known. It is of course possible to introduce dissipation and/or adjust reflection coefficients or breakwater closure to tune the model better so that a calibrated model for the Los Angeles/Long Beach complex would be available for future use. However, there is no assurance that the hydraulic model is the true benchmark. The high level of agreement between the hydraulic model and numerical model results for the short and medium period waves and the moderate agreement for the long period waves indicates that the performance of the numerical and hydraulic models are certainly compatible, although not identical.
154
V. Panchang & Z.
Demirbilek
Wave Periods from 30s to 42s
— Model -Q-Data, Upper Bound -<>-Data, Lower Bound
5
10
15
20
25
30
35
40
45
50
55
45
50
55
50
55
Wave Periods from 42s to 205s
5
10
15
20
25
30
35
40
Wave Periods from 205s to 512s
10
15
20 25 30 35 Gage Number
40
45
Fig. 11. Wave height comparison for Los Angeles/Long Beach Harbor complex.
Simulation
5.2. Simulations
in Barber's
Point
of Waves in Harbors
155
Harbor
Seiches in Barber's Point Harbor (Hawaii) have been studied by Okihiro et al. (1993) and Okihiro and Guza (1996). Okihiro et al. (1993) also used a numerical model (Chen and Houston, 1987) based on Eq. (5) with Eqs. (15)(17) describing the open boundary conditions. Such a model, as noted earlier, confronts the modeler with having to select a constant exterior depth and to assume that the exterior coastline is fully reflecting. Further, this model solves Eq. (28) by Gausian elimination. As noted earlier, this creates storage problems and hence allows only coarse resolution for some frequencies. To overcome these limitations, Eqs. (18), (20) and (27) were used to formulate the open boundary conditions. Bathymetric data used by Okihiro et al. (1993) were used to develop a new grid containing about 65,065 elements. The model was run for 136 frequency components. Full reflection was used on all closed boundaries. Field data were available at four locations inside the harbor (denoted by East, West, North and South gages); see Fig. 12. Data were also available at a gage outside the harbor (denoted by "offshore" gage in Fig. 12); these were used to normalize the amplification factors inside the harbor. Model results are
Fig. 12. Barber's Point Harbor, bathymetry and gage locations.
156
V. Panchang & Z.
Demirbilek
south
+ model - observations 0.005
0.005
0.01
0.015 north
0.02
0.025
0.03
0.01
0.015
0.02
0.025
0.03
0.015 0.02 frequency (Hz)
0.025
0.03
west
0.005
0.01
Fig. 13. Wave height comparison for Barbers Point Harbor.
compared against field data in Fig. 13. There is fairly good agreement between the model calculations and the measurements especially for the long periods. For the short periods, there appears to be some overprediction by the model. This is attributed to the fact that shorter waves experience less reflection. The
Simulation
of Waves in Harbors
157
simulations with a lower reflection coefficient for these waves and more detailed results will be presented elsewhere. However, the results at all four locations inside the harbor are a fairly reasonable reproduction of the field data. 6. Concluding Remarks In this paper, we have provided a review of recent developments in simulating ocean waves with models based on the elliptic refraction-diffraction equation. In general, finite element models appear to be best suited for practical applications covering the full spectrum of waves to which a harbor may be exposed. (Some practical applications may be found, for example, in Tang et al., 1999; Pos et al, 1989; Mattioli, 1996; Kostense et al., 1988). Advances in the treatment of boundary conditions and of matrix systems associated with the discretized equations have made it possible to eliminate many of the difficulties that led to inferior solutions. They have also eliminated the need for approximations of the elliptic model. Further, the advances reduce the burden on the modeler who does not have to test the sensitivity of model results to unrealistic assumptions (such as constant depths in the exterior). Applications to the Los Angeles/Long Beach Harbor region and to Barber's Point Harbor presented here demonstrate that finite element modeling with the techniques described in this paper produces results that are at least as reliable as those obtained by other methods. Inclusion of additional mechanisms like dissipation, wave-wave interactions, wave-current interactions and steep slope effects can enhance the usefulness of these models. However, how a model will behave when these effects are combined is not yet clear. Further research in modeling methods as well as data where some of these effects can be combined and isolated are desirable. Acknowledgments Partial support for this work was provided by the Office of Naval Research, the Maine Sea Grant Program and the National Sea Grant Office. Many of the developments described here were made with the assistance of five graduate students at the University of Maine (Karl Schlenker, Wei Chen, Luizhi Zhao, Doncheng Li and Khalid Zubier) and Dr. Michele Okihiro of the Scripps Institution of Oceanography. Their contributions are gratefully acknowledged. Permission to publish this paper was granted by the Chief, Corps of Engineers, to publish this paper.
158
V. Panchang & Z. Demirbilek
References Battjes, J. A. and J. Janssen (1978). Energy loss and set-up due to breaking of random waves. Proc. 16th Int. Conf. Coastal Eng., ASCE, New York, 569-587. Berkhoff, J. C. W. (1976). Mathematical Models for Simple Harmonic Linear Water Waves. Wave Refraction and Diffraction, Publ. 163, Delft Hydraulics Laboratory. Beltrami, G. M., G. Bellotti, P. De Girolamo, and P. Sammarco (2000). Treatment of wave breaking and total absorption in a mild-slope equation FEM model. J. Waterway, Port, Coastal & Ocean Eng. To appear. Booij, N. (1981). Gravity Waves on Water with Nonuniform Depth and Current, Ph.D. Thesis, Technical Univ. of Delft, The Netherlands. Bova, S. W., C. P. Breshears, C. Cuicchi, Z. Demirbilek, and H. A. Gabb (2000). Dual-level parallel analysis of harbor wave response using MPI and OpenMPI. Int. J. High Performance Comput. Appl. 14(1): 49-64. Chamberlain, P. G. and D. Porter (1995). The modified mild-slope equation. J. Fluid Mech. 291: 393-407. Chandrasekera, C. N. and K. F. Cheung (1997). Extended linear refraction-diffraction model. J. Waterway, Port, Coastal & Ocean Eng. 123(5): 280-286. Chawla, A., H. T. Ozkan-Haller, and J. T. Kirby (1998). Spectral model for wave transformation and breaking over irregular bathymetry. J. Waterway, Port, Coastal & Ocean Eng. 124: 189-198. Chen, H. S. (1986). Effects of bottom friction and boundary absorption on water wave scattering. Appl. Ocean Res. 82(2): 99-104. Chen, H. S. (1990). Infinite elements for water wave radiation and scattering. Int. J. Numer. Meth. Fluids 11: 55-569. Chen, H. S. and J. R. Houston (1987). Calculation of water level oscillations in harbors. Instructional Rept. CERC-87-2, Waterways Expt. Stn., Vicksburg, Mississippi. Chiang, W.-L. (1988). Modeling long and intermediate waves in a harbor. Appl. Math. Modeling 12: 423-428. Dally, W. R., R. G. Dean, and R. A. Dalrymple (1985). Wave height variation across beaches of arbitrary profile. J. Geophys. Res. 90(c6): 11917-11927. Dalrymple, R. A., J. T. Kirby, and P. A. Hwang (1984). Wave diffraction due to areas of high energy dissipation. J. Waterway, Port, Coastal & Ocean Eng. 110(1): 67-79. Demirbilek, Z., B. Xu, and V. G. Panchang (1996). Uncertainties in the validation of harbor wave models. Proc. 25th Int. Coastal Eng. Conf., 1256-1267. Demirbilek, Z. and V. G. Panchang (1998). CGWAVE: A coastal surface water wave model of the mild slope equation, Tech. Rept. CHL-98-26, US Army Corps, of Engineers Waterways Expt. Stn., Vicksburg, MS 39180. Dickson, W. S., T. H. C. Herbers, and E. B. Thornton (1995). Wave reflection from breakwater. J. Waterway, Port, Coastal & Ocean Eng. 121(5): 262-268.
Simulation
of Waves in Harbors
159
Dingemaans, M. W. (1997). Water Wave Propagation over Uneven Bottoms. World Scientific. Singapore. Ebersole, B. A. (1985). Refraction-diffraction model for linear water wave. J. Waterway, Port, Coastal & Ocean Eng. 111(6): 939-953. Foster, I. (1997). Designing and Building Parallel Programs, Addison-Wesley Publ. Co., Reading, MA. Givoli, D. (1991). Nonreflecting Boundary Conditions. J. Comput. Phys. 94: 1-29. Hedges, T. S. and B. W. Lee (1992). The equivalent uniform current in wave-current interaction computations. Coastal Eng. 16: 301-311. Hurdle, D. P., J. K. Kostense, and P. Bosch (1989). Mild slope model for the wave behavior in and around harbors and coastal structures. In: Advance in Water Modeling and Measurement, ed. M. H. Palmer. BHRA, The Fluid Eng. Centre, Cranfield, England. 307-324. Irons, P. (1970). A frontal solution program for finite element analysis. Int. J. Numer. Meth. Eng. 2: 5-32. Isaacson, M. and S. Qu (1990). Waves in a harbor with partially reflecting boundaries. Coastal Eng. 14: 193-214. Isaacson, M., E. O'Sullivan, and J. Baldwin (1993). Reflection effects on wave field within a harbor. Can. J. Civ. Eng. 20(3): 386-397. Isobe, M. (1999). Equation for numerical modeling of wave transformation in shallow water. In: Developments in Offshore Engineering, Chapter 3, ed. J. B. Herbich. Gulf Publishing, Houston. 101-162. Jeong, W. M., J. W. Chae, W. S. Park, and K. T. Jung (1996). Field measurements and numerical modeling of harbor oscillations during storm waves. Proc. 25th Int. Conf. Coastal Eng. ASCE, New York. 1268-1279. Jones, N. L. and D. R. Richards (1992). Mesh generation for estuarine flow models. J. Waterway, Port, Coastal & Ocean Eng. 118(6). Kamphuis, J. W. (1994). Wave height from deep water through breaking zone. J. Waterway, Port, Coastal & Ocean Eng. 120(4): 347-367. Kaihatu, J. and J. T. Kirby (1995). Nonlinear transformation of waves in finite water depth. Phys. Fluids 7(8): 1903-1914. Kirby, J. T. (1984). A note on linear surface wave-current interaction over slowly varying topography waves. J. Geophys. Res. 8 9 ( d ) : 745-747. Kirby, J. T. (1986). Higher order approximation in the parabolic equation method for water waves. J. Geophys. Res. 9 1 ( d ) : 933-952. Kostense, J. K., K. L. Meijer, M. W. Dingemans, A. E. Mynett, and P. van den Bosch (1986). Wave energy dissipation in arbitrarily shaped harbors of variable depth. Proc. 20th Int. Conf. Coastal Eng. 2002-2016. Kostense, J. K., M. W. Dingemans, and P. van den Bosch (1988). Wave-current interaction in harbors. Proc. 21th Int. Conf. Coastal Eng. 1: 32-46. ASCE, New York. Lee, J. J. and F. Raichlen (1972). Oscillations in harbors with connected basins. J. Waterways, Harbors, and Coastal Eng. Div. 98: 311-332. ASCE.
160
V. Panchang & Z.
Demirbilek
Lennon, G. P., P. L.-F. Liu, and J. A. Liggett (1982). Boundary integral solutions of water wave problems. J. Hydr. Div. ASCE 108: 921-931. Li, B. (1994a). A generalized conjugate gradient model for the mild slope equation. Coastal Eng. 23: 215-225. Li, B. (1994b). An evolution equation for water waves. Coastal Eng. 23: 227-242. Li, B. and K. Anastasiou (1992). Efficient elliptic solvers for the mild-slope equation using the multigrid method. Coastal Eng. 16: 245-266. Li, B., D. E. Reeve, and C. A. Fleming (1993). Numerical solution of the elliptic mild-slope equation for irregular wave propagation. Coastal Eng. 20: 85-100. Larson, M. (1995). Model for decay of random waves in surf zone. J. Waterway, Port, Coastal & Ocean Eng. 121(1): 1-12. Massel, S. R. (1992). Inclusion of wave-breaking mechanism in a modified mild-slope model. In: Breaking Waves IUTAM Symposium, Sydney/Australia, 1991. eds. M. L. Banner and R. H. J. Grimshaw. Springer-Verlag, Berlin Heidelberg, 1992. 319-324. Massel, S. R. (1993). Extended refraction-diffraction equation for surface waves. Coastal Eng. 19: 97-126. Massel, S. R. and M. R. Gourlay (2000). On the modeling of wave breaking and set-up on coral reefs. Coastal Eng. 391: 1-27. Marchuk, G. I. (1975). Methods of Numerical Mathematics. Applications of Mathematics. Springer-Verlag, New York. Mattioli, F. (1996). Dynamic response of the Lido channel to wave motion in the presence of movable barriers. // Nuovo Cimento 19c(l): 177-194. Mei, C. C. (1983). The Applied Dynamics of Ocean Surface Waves. John Wiley, New York. Moffatt k. Nichol Engineers (1999). San Pedro Bay Harbor Resonance Model for LA/LB Complex. User's Manual. 250 W. Wardlow Road, Long Beach, CA 90807. OARB (1997). OpenMP Fortran Application Program Interface. OpenMP Architecture Review Board (OARB) vl.0, http://www.openmp.org, October 1997. Okihiro, M. and R. T. Guza (1996). Observations of Seiche forcing and amplification in three small harbors. J. Waterway, Port, Coastal & Ocean Eng. 122(5): 232-238. Okihiro, M., R. T. Guza, and R. J. Seymour (1993). Excitation of Seiche observed in a small harbor. J. Geophys. Res. 122(5): 232-238. Oliveira, F. S. B. F. and K. Anastasiou (1998). An efficient computational model for water wave propagation in coastal regions. Appl. Ocean Res. 20: 263-271. Panchang, V. G., B. Cushman-Roisin, and B. R. Pearce (1988). Combined refractiondiffraction of short waves for large coastal regions. Coastal Eng. 12: 133-156. Panchang, V. G., W. Ge, B. Cushman-Roisin, and B. R. Pearce (1991). Solution to the mild-slope wave problem by iteration. Appl. Ocean Res. 13(4): 187-199. Panchang, V. G., B. Xu, and B. Cushman-Roisin (1993). Bathymetric variations in the exterior domain of a harbor wave model. Proc. Int. Conf. Hydroscience and Eng. Washington DC. 1555-1562.
Simulation
of Waves in Harbors
161
Panchang, V. G., B. Xu, and Z. Demirbilek (1999). Wave prediction models for coastal engineering applications. In: Developments in Offshore Engineering, ed. J. B. Herbich. Chapter 4. Gulf Publishing, Houston. 163-194. Panchang, V. G., W. Chen, B. Xu, K. Schlenker, Z. Demirbilek, and M. Okihiro (2000). Effects of exterior bathymetry in elliptic harbor wave models. J. Waterway, Port, Coastal & Ocean Eng. 126(2): 71-78. Pos, J. D. (1985). Asymmetrical breakwater gap wave diffraction using finite and infinite elements. Coastal Eng. 9: 101-1123. Pos, J. D. and F. A. Kilner (1987). Breakwater gap wave diffraction: An experimental and numerical study. J. Waterway, Port, Coastal & Ocean Eng. 113(1): 1-21. Pos, J. D., J. W. Gonsalves, and A. H. Holtzhausen (1989). Short-wave penetration of harbors: A case study at Mossel Bay. Proc. 9th Ann. Conf. Finite Element Meth., February 8-10, Stellenbosch, South Africa. Porter, D. and D. J. Staziker (1995). Extensions of the mild-slope equation. J. Fluid Mech. 300: 367-382. Radder, A. C. (1979). On the parabolic equation method for water-wave propagation. J. Fluid Mech. 95: 159-176. Radder A. C. (1992). Efficient elliptic solvers for the mild-slope equation using the multigrid method. Coastal Eng. 18: 347-352. Schaffer, H. A. and I. G. Jonsson (1992). Edge waves revisited. Coastal Eng. 16: 349-368. Seabergh, W. C. and L. J. Thomas (1995). Los Angeles Harbor Pier 400 Harbor Resonance Model Study. US Army Corps, of Engineers Waterways Expt. Stn., Vicksburg, MS 39180. TR CERC-95-8. Smith, G. D. (1978). Numerical Solution of Partial Differential Equations: Finite Difference Mathods. Oxford University Press. Smith, R. and T. Sprinks (1975). Scattering of surface waves by a conical island, J. Fluid Mech. 72: 373. Steward, D. R. and V. G. Panchang (2000). Improved coastal boundary conditions for water wave simulation models. Ocean Eng. 28: 139-157. Sutherland, J. and T. O'Donoghue (1998). Wave phase shift at coastal structures. J. Waterway, Port, Coastal & Ocean Eng. 124(2): 80-98 Tang, Y. and Y. Ouellet (1997). A new kind of nonlinear mild-slope equation for combined refraction-diffraction of multifrequency waves. Coast. Eng. 3 1 : 3-36. Tang, Y., Y. Ouellet, and Y. Ropars (1999). Finite element modeling of wave conditions inside Sainte-Therese-de-gaspe Harbor, Quebec. Proc. Canadian Coastal Conf. 737-748. Thompson, E. F., H. S. Chen, and L. L. Hadley (1996). Validation of numerical model for wind waves and swell in harbors. J. Waterway, Port, Coastal & Ocean Eng. 122(5): 245-256. Tsay, T.-K. and P. L.-F. Liu (1983). A finite element model for wave refraction and diffraction. Appl. Ocean Res. 5(1): 30-37.
162
V. Panchang & Z. Demirbilek
Tsay, T.-K., W. Zhu, and P. L.-F. Liu (1989). A finite-element model for wave refraction, diffraction, reflection, and dissipation. Appl. Res. 11: 33-38. Whalin, R. W. (1971). The Limit of Application of Linear Refraction Theory in A Convergence Zone. US Army Corps, of Engineers Waterways Experiment Station, Vicksburg. Research Rept H-71-3. Xu, B. and V. G. Panchang (1993). Outgoing boundary conditions for elliptic water wave models. Proc. R. Soc. London, Ser. A 4 4 1 : 575-588. Xu, B., V. G. Panchang, and Z. Demirbilek (1996). Exterior reflections in elliptic harbor wave models. J. Waterway, Port, Coastal & Ocean Eng. 122(3): 118-126. YOTO (1998). "Our Ocean Future", Year of the Ocean, Themes & Issues Concerning the Nation's Stake in the Oceans. Prepared by the H. John Heinz III Center for Science, Economics, and the Environment. Office of the Chief Scientist, NOAA, Washington, DC 20230. Zhao, L., V. G. Panchang, W. Chen, Z. Demirbilek, and N. Chhabbra (2000). Simulation of breaking effects in a two-dimensional harbor wave prediction model. Coastal Eng. To appear. Zundell, A. K., A. L. Fugal, N. L. Jones, and Z. Demirbilek (1998). Automatic definition of two-dimensional coastal finite element domains. In: Hydroinformatics98, Proc. 3rd Int. Conf. Hydroinformatics. eds. V. Babovic and L. C. Larsen. A. A. Balkema, Rotterdam. 693-700.
R E C E N T A D V A N C E S IN T H E MODELING OF WAVE A N D PERMEABLE STRUCTURE INTERACTION
INIGO J. LOS ADA Artificial and natural porous structures are of great interest in coastal and harbor engineering. The modeling of wave interaction with permeable structures is therefore a key issue to determine the functionality and stability of this kind of structures. In most circumstances, an averaging process is introduced in the analysis of the flow in terms of a seepage or discharge velocity and some coefficients depending on the flow. In order to solve the wave and structure interaction, the porous flow model is matched with a flow model for the fluid region. In this paper, it will be shown that several new equations including the resistance forces in the porous medium have been derived. Newly developed models based on Boussinesqtype equations or direct resolution of the Navier—Stokes equations using VOF techniques have opened a new range of possible applications. However, these models still highly depend on porous flow coefficients. Predictive formulae for these constants under oscillatory flow conditions require further research especially if these models are considered to be an alternative to physical modeling in the design of coastal structures.
1. Introduction A porous medium is a two-phase material in which the solid matrix, usually assumed to be rigid, constitutes one phase and the interconnected voids or pores constitutes the other. One of the main characteristics of porous media is the irregular shape and size of its pores, randomly distributed, conferring the flow through this heterogeneous formation considered a very complex nature. Our interest will be to determine the flow through the porous formation with typical length scales much larger than the characteristic pore size. Artificial porous structures such as rubble-mound breakwaters, submerged structures, outfall protections, artificial fishing reefs or armor layers for the protection of seawalls or vertical structures are of great interest in coastal and harbor engineering since they provide one of the best means to induce incident wave dissipation by friction inside the structures. Therefore, the knowledge 163
164
/. J. Losada
of the flow motion in and around the porous media and the corresponding pressure and force fields will be a key issue to determine the functionality and stability of these coastal structures. Furthermore, sand and shingle beaches may be considered as the n a t u r a l porous media and therefore the determination of wave dissipation on permeable layers, the beach water table evaluation, or of how beach permeability may affect the fundamental surf zone hydrodynamic processes are also important issues to be addressed. T h e complex internal geometry of a porous medium, artificial or natural, is difficult if not impossible to determine. Furthermore, in general, in the coastal and harbor engineering field, there is a lack of interest in knowing the internal details of the structure or the microscopic flow. In fact, in most circumstances, our interest will be in determining the characteristics of the flow in large portions of the porous structure considered and introducing an averaging process in the analysis of the flow. T h e averaging process has a smoothing effect, filtering out small scale variations associated with t h e media heterogeneity and pore irregularities. In several other fields, the transition to average macroscopic variables is based on a statistical approach, especially in hydrology when analyzing flow and t r a n s p o r t in porous formations and aquifers. However, to date, the stochastic approach is not considered yet in the field of coastal engineering since several additional complications have to be addressed. T h e modeling of wave and permeable n a t u r a l or artificial structure interaction is based on the coupling of two models, one t h a t describes the flow acting on the structure and one t h a t describes the flow through the porous structures. T h e accuracy of the modeling will be limited by the hypotheses and simplifications formulated for the flow in the outer fluid region a n d by the validity and hypotheses of the porous flow model, usually relying on some constants depending on the flow and finally on the matching conditions imposed. This paper summarizes some of the most recent work available in the literature on wave interaction with porous structures. It will be shown t h a t the study of wave interaction with permeable structures has evolved in parallel with wave theories in fluids. In the last few years, special attention has been paid to the development of equations and numerical models to analyze the interaction with porous structures with very promising results. However, the application of these models to prototype scale has to be carried out with care. This paper is organized as follows. First, a review of the existing porous flow models is presented. An emphasis is made on the difference between
Recent Advances in the Modeling of Wave
165
stationary and nonstationary flows. In Sec. 3, the general governing equations for flow in porous media are derived. In order to formulate simple solutions, the linearized problem is shown in Sec. 4. Considering potential flow inside and outside the porous medium, solutions in terms of eigenfunction expansions are formulated to analyze wave interaction with vertical permeable structures. Furthermore, the derivation of an extended mild-slope equation for wave propagation on permeable layers is presented. The equation is applied to wave interaction with submerged permeable breakwaters including wave breaking. In Sec. 5, the most recent developments for shallow water equations are indicated. Wave diffraction and transmission by a permeable vertical breakwater is modeled using Boussinesq-type equations. The time and depth-averaged equations for waves in permeable media are derived in Sec. 6. Some preliminary applications are shown to model mean water level variations in permeable submerged breakwaters. Finally, in Sec. 7, a model based on the Navier-Stokes equation is presented. This model, called COBRAS is able to simulate wave interaction with permeable structures including wave breaking and turbulence. 2. Porous Flow Models 2.1. Stationary
flow
The success of the theoretical formulation of the wave and porous structure interaction largely depends on the accuracy of the empirical formulas and coefficients used to describe the frictional forces exerted by the porous media. The study of the flow in porous media can be traced back to 1856 when Darcy found empirically, using a vertical permeameter, that the ID steady flow in sand or other fine granular material can be described by the following formula (Bear, 1972), I = K-lud = apud , (1) where i" = (—dpo/dy)/pg is the hydraulic gradient, p the fluid density, g the gravitational acceleration, po = p + pgho the effective pressure, ho the vertical distance from the selected datum, ud the discharge velocity and K = 1/a.p (m/s) the permeability coefficient and ap an empirical coefficient. The Darcy law is often written in the form I = (v/gKp)ud where Kp (m 2 ) is the intrinsic permeability. This expression, which later has been referred to as the Darcy law, is esentially valid for laminar flows, breaking down when the flow velocity becomes sufficiently large or when the characteristic length scale of the porous material is large. A certain critical Reynolds number
166
/. J. Losada
Re = ucDc/v above which Darcy law becomes invalid can be defined in terms of a characteristic discharge velocity in the porous media uC: a characteristic length scale of porous media Dc and the molecular viscosity v. Above this critical number, the turbulence effect becomes apparent and the flow resistance appears to increase. Forchheimer (see Bear, 1972), at the beginning of the century extended Darcy's law to include a quadratic term accounting for the frictional force induced by turbulence such that: I = apud + bp\ud\ud,
(2)
where bp is another empirical coefficient with dimension (s 2 /m 2 ). In order to describe the coefficients ap and bp, several expressions were proposed. Some of them were obtained based on various analogies. Considering nonconvective laminar flow in a number of circular capillaries with diameter dc, Darcy's law compares to the Hagen-Poiseuille equation yielding to the following relation, 7= 3 2 ^ . (3) Applying the pipe analogy, the combined efforts of Kozeny (1927) and Carman (1937) lead to: / = 3 6 « < ^ - ^ « ' , n%
(4)
gdz
where ne is the porosity defined as the fluid volume divided by the total volume d is the diameter of spheres with equivalent geometry to the pores and K is a coefficient which is taken to be 0.5. Ergun (1952) extended the work by Kozeny and Carman for the linear flow resistance applying the pipe analogy to the Forchheimer flow regime and introducing a new constant for the quadratic term arriving at the following expression, / = 150 { i z £ > ! » n%
gd1
u
' + 1.75 i i Z ^ l n%
VI-
(5)
gd
In general, the following expressions for the ap and bp coefficients can be used, (l-ne)2 v ap=iap
^^g^!>
(6)
Recent Advances in the Modeling of Wave
6P = / 3 P ^ 4 r n%
167
(7)
gDc
A summary of the different values of the nondimensional coefficients ap, j3p and the characteristic length-scale for the porous media Dc proposed by several authors can be found in van Gent (1993). 2.2. Nonstationary
flow
Wave action on structures induces a nonstationary flow. Polubarinova-Kochina (1962) added a time-dependent term to the Forchheimer equation as Eq. (2). The resulting equation including the inertia term accounting for the acceleration is referred to as the extended Forchheimer equation, dud I = apud + bpud\ud\ + cp— ,
(8)
where cp is a dimensional coefficient (s 2 /m). This formula can be derived from the Navier-Stokes equation (van Gent, 1991). Gu and Wang (1991) and van Gent (1991) found an expression for cp after a theoretical derivation, cP =
- ^ ~ , (9) neg where 7 is a nondimensional coefficient that accounts for the added mass. The concept of added mass is associated with the fact that in order to accelerate a certain volume of water, a certain amount of momentum is needed. To accelerate the same volume of water in a porous medium, an additional amount of momentum is needed. This is called added mass since the extra amount of momentum suggests that a larger volume of fluid has to be accelerated. Please note that Eq. (8) does not include a possible resistance force due to the presence of a convective term. Such a resistance term, probably important for flow through porous media with considerable large-scale convective transport, could be incorporated into the bp term because it would be quadratic in the velocity (van Gent, 1991). 2.2.1. Relative importance of the resistance forces Gu and Wang (1991), van Gent (1993) and Losada et al. (1995) discussed the relative importance of the three contributions to the total resistance force for nonstationary flow in porous media represented in Eq. (8), namely the
168
/. J. Losada
resistance due to laminar flow f^, resistance due to turbulent flow /jy, and the inertial resistance / / . Following Gu and Wang (1991), the importance of the resistance forces can be analyzed in terms of two different Reynolds-type numbers, D
Kf =
ucDc
,
<10)
a-^-,3. Kl
~ v ~l KC' where uc is a characteristic velocity, Dc is the characteristic particle size of the porous material, v is the kinematic viscosity and KC is the Keulegan-Carpenter number and to is the angular wave frequency. Both Reynolds numbers give the relative importance of the inertial forces to the viscous forces; however, in Rf, the inertia is of convective nature and the resistance is due to velocity changes in space, whereas in Ri, the resistance force arises locally due to the change of velocity at a specific location of the porous structure. In Fig. 1, regions with different dominant resistance components are shown. Regions / , L, and N are dominated by one resistance force only, that is, the dominant force is at least one order of magnitude larger than the other two. 1E6
fN > 10fL fn > 10f, N Region
1E5 IE4 1E3
h > JN
1E2
1 El
k > I Of, h > 10fH L Region
Q i ~ 1E0
1 El
f,>10fL f, > 10fH /Region
1E-2 1E-3
1E-5
Smith (1991) • - - vanGent(1994) Losada et al. (1995)
1E-6 IE-6
1E-S
1E-4
1E-4
1E-3
IE-2
1 El
1E0
A./, 1 El
1E2
1E3
1E4
1E5
1E6
Fig. 1. Relative importance of the resistance forces (after Gu and Wang, 1991).
Recent Advances in the Modeling of Wave
169
Table 1. Dominant force components under coastal wave actions (Gu and Wang, 1991). Material description Coarse sand or finer Pebble or small gravel Large gravel Crushed stone Boulders Crushed stone Artifical blocks Large rocks
D(m)
uc (m/s)
Rf
Rt
< 0.002
< O(10~ 3 )
< O(10°)
< O(10°)
0.01
O(10"2)
O(10 2 )
O(10 2 )
0.10
O(10_1)
O(10 4 )
O(10 4 )
0.3-1.0
O(10°)
O(10 6 )
O(10 5 )
> 1.0
> O(10°)
> O(10 6 )
> O(io6)
Dominant force Laminar Laminar Turbulence Inertia Turbulence Inertia Turbulence Inertia Turbulence Inertia
In one region, the three resistance forces are of equal importance while there are three intermediate regions where two out of the three forces could be important. Considering several characteristic parameters under coastal wave conditions, Gu and Wang (1991) gives an illustration of dominant force components for bottom material of various sizes. This table is only orientative but it is a general guideline of practical interest. Furthermore, the results in Fig. 1 and Table 1 are important for analyzing possible scale effects in physical models. 2.2.2. Determination of the parameters for nonstationary flow The friction coefficients ap and bp from the Forchheimer equations were measured in tests with stationary flow. Until the work by Smith (1991), Hall et al. (1995) and van Gent (1993, 1995), no available data for the determination of the friction coefficients under oscillatory flow were reported. Smith (1991) provided a set of friction coefficients obtained experimentally in an oscillating water tunnel through different arrangements of prepared packing spheres. One sample of rock material was tested. van Gent (1993, 1995) carried out permeability measurements in a U-tube tunnel to study flow through five samples with various types of stones with D50 = 0.0202 — 0.0610 m. The differences between stationary and oscillatory flow were studied and the contributions of laminar, turbulence and inertia
170
/. J. Losada
terms were determined. Figure 1 shows the range where the experimental work by Smith (1991), van Gent (1993) and Losada et al. (1995) were carried out. Based on the experimental results, new expressions for nonstationary porous flow friction coefficients were formulated. According to van Gent (1995), the friction coefficients for the extended Forchheimer equation should be expressed as: (l-ne)2
v
gDl '50 ' Ml v
\
T+
i ^ ^ - k KCJ n\ gD50
1 + 7 ^ , „„ — where 7 = 0.85 neg uc 0.015 negT > j ^ + 0.85 '
where KG 0.015 — Ac
^ ne-^50 eD
(11)
where
where uc is a characteristic velocity of the flow, T is the wave period and D$Q is the median grain size diameter. It is recommended to take the maximum discharge velocity as representative of the flow. Although the coefficients ap and j3p may still depend on parameters like grading, shape, aspect ratio or orientation of the stones, the following values are recommended, ap = 1000 and j3p = 1.1. Further work on unsteady flow equations can be found in Burcharth and Andersen (1995). Equation (11) represent a significant contribution to the modeling of wave interaction with permeable structures. However, the discrepancies appeared in the application of the formulae by other authors (see, Liu et al., 1999; Lynett et al., 2000); the limited range of existing experimental data and the importance of an accurate prediction of the coefficients on the modeling of wave and structure interaction seem to be important reasons to do further research on this topic in the near future especially if the modeling is to be applied to prototypes. 3. General Governing Equations and Matching Conditions 3.1. Governing
equations
The flow in porous media can be described by the general Navier-Stokes equations,
Recent Advances in the Modeling of Wave
du*
* ^ i _ _ l ^
a2<
171
s
where v is the molecular viscosity, u* is the ith component of the instantaneous velocity in the pores and p* the instantaneous effective pressure. As already stated using the macroscopic approach, based on averaging over a small but finite volume with a representative length scale larger than the typical pore size but smaller than the characteristic length scale of the problem, perturbation in the field due to the presence of individual particles and pore irregularities can be ignored. To replace the actual velocity with the seepage velocity, Sollitt and Cross (1972) resolved the local instantaneous velocity field u* into three components, u* = Ui + u\ + u\,
(14)
where Ui is the seepage velocity, that is, "the average velocity within small but finite and uniformly distributed void spaces; u\ is the spatial perturbation accounting for local velocity components due to pore irregularities or boundary layers, and u\ is the time perturbation accounting for local transient fluctuations within the pores" (Sollitt and Cross, 1972). Likewise, the pressure field may be split up into analogous components. The effect of the transient or turbulent and spatial perturbations on the mean flow in the pore can be determined by substituting these definitions in the Navier-Stokes equations. Expanding the total derivative in the NavierStokes equations, substituting Eq. (14) and the analogous expression for the pressure field in this equation, and performing the time averaging for a period much smaller than the time scale of the macroscopic unsteadiness yields, — (Ui + u\) + {ut + u\) • V(Ui + u\) + = — <j> + p"+iz)+vV2(ui P
vfiv\
+ u\),
(15)
where (A) denotes time average in a period of time smaller than the wave period. Proceeding in the same way in the continuity equations leads to: V - ( U i + u ? ) = 0.
(16)
172
I. J. Losada
Integrating the equations of motion over a small but finite volume, the effect of spatial fluctuations within the pore may be isolated. Finally, the following set of Reynolds Averaged Navier-Stokes equations (RANS) are obtained, + < • VM| = — V(p + 7-z) + vV2Ui
- ^ + UiVui + uyVut
(17a)
and V«i = 0 ,
(17b)
where (A) denotes spatial average. Due to the nonlinearity related to the convective terms, the terms associated to the spatial and turbulent fluctuations uf • Vuf and u\ • Vu* remain in the equations after the averaging process. In analogy to turbulence analysis, these two terms may be interpreted as stresses with respect to the mean motion. In order to solve the equations, it is necessary to find closure equations. Based on the work by Ward (1964) for steady and nonconvective flow conditions, Sollitt and Cross (1972) established the following equivalency, vV2Ui - [ u p V u ^ + u f - V u * ] = - vul
u-t\uf\
(18)
where Kp is the intrinsic permeability, Cf a dimensionless turbulent coefficient and uf the discharge velocity which is related to the seepage velocity by the following relationship uf = neUi, where ne is the porosity. The following equation is obtained, dui
1
.
vneUi
Crrr
(19)
where s = necpg is introduced as a co-factor in the local acceleration term to account for the added mass and the following additional assumption has been made, UjVuj < usj • Vut + uyVul.
(20)
Sollitt and Cross (1972) claimed this assumption to be valid for problems of practical importance, for which wave length is much greater than pore diameter. Please note that the viscous terms in Eq. (17a) have been included in
Recent Advances in the Modeling of Wave
173
the equivalency. One could have assumed this term to be negligible compared to the fluctuations. Using a vertical-lattice type porous medium made of rectangular wooden sticks nailed together producing uniform pores, Losada et al. (1995) performed a very comprehensive set of measurements inside and outside the structure including free surface and water particle velocities using Laser Doppler Velocimetry and pressure records. Instantaneous measurements at 9 points at each pore were processed in order to calculate the instantaneous and spatial fluctuations, the time and space-averaged velocity, and the convective terms associated with each of the components. The analysis points out two important results: (1) spatial fluctuations are always more important than temporal fluctuations since temporal fluctuations are confined by the pore size, and (2) even if under certain circumstances, convective terms associated with fluctuations are more important than those associated with seepage flow, the latter should not be neglected. There is no unique way to decompose the instantaneous velocity. Liu et al. (1999, 2000) supported by the experimental results by Losada et al. (1995) considered dividing the fluid variables into two parts only, a spatially-averaged component and a spatially-fluctuating component, assuming the temporal fluctuations to be negligible. The resulting spatially-averaged Navier-Stokes equations are: 1 + cA duf ne dt
ufduf_ nl dxj
I dp p dxt
v d2uf ne dxjdxj
1 duiuj n% dxj '
(21)
where CA = 7((1 — ne)/ne) is the added mass coefficient. The correlation of spatial velocity fluctuations, the last term on the righthand side of Eq. (21) is modeled by Liu et al. (1999) by a combination of linear and nonlinear frictional forces as follows, I duluSj 2
TT~
= ~gaPUi
- gbpucUi,
(22)
in which uc = y/uiul so that the first term on the right side represents the linear, viscous force, while the second term represents the nonlinear turbulent force. Both ap and bp are empirical coefficients which are functions of Reynolds number and the geometric characteristics of the porous media.
174
/. J. Losada
Please note that Eqs. (22) and (18) are similar. However, the second (viscous) term on the right-hand side in Eq. (17), generally much smaller than the third term for problems of engineering interest, is retained in Liu et al. (1999) since this term is responsible for transferring shear force and may become increasingly important near the interface between porous media and outside flow for smaller scale problems. 3.2. Matching
conditions
Matching conditions are necessary to guarantee the continuity of the solution for the interface between the fluid and porous regions. In general, continuity of mass flux and pressure are the matching conditions considered. For long wave models, continuity of free surface, velocity and their derivatives are usually enforced. However, when the models in the fluid region or inside the porous structure include the modeling of turbulence or boundary layers, the matching at the interface has to be carried out with caution. 4. Wave Interaction with Structures. Linear Solutions 4.1. Linearized
problem
Assuming a simple harmonic wave of frequency to, Eq. (19) may be linearized on the basis of Lorentz's hypothesis of equivalent work (Sollitt and Cross, 1972; Madsen, 1974), replacing the nonlinear terms in Eq. (19) by an equivalent linear term fuJUi where / is a dimensionless friction coefficient. This yields a linearized form of the equation, istoui — —V • I —h gz ) — tofui.
\P
(23)
J
Taking the curl of this equation shows that the flow in the porous medium is irrotational and therefore can be described by a potential $ that satisfies, m = V$ .
(24)
Substituting Eq. (24) into Eq. (23) results in a Bernoulli-type equation for unsteady flow within the porous medium, s^-
+ -+gz
+ fuj<S> = 0.
(25)
Recent Advances in the Modeling of Wave 175
Finally, substituting Eq. (24) into the continuity equation in Eq. (17b) yields Laplace's equation, V 2 $ = 0. (26) At the free surface 77, the Bernoulli equation in Eq. (25) can be combined with the linear kinematic free surface boundary condition,
to yield, ?l-u>2(s-if)-=0. (28) dz g Furthermore, the following linear complex dispersion relationship has to be satisfied by the waves propagating inside the porous medium, to2(s-if)=
gTtanhTh,
(29)
where T is a complex wavenumber. Solutions to these equations depend on the values of the porous material parameters, s, ne, Kp, and Cf, known for a given material and the linearized friction coefficient / . Therefore, an additional condition is required to evaluate this coefficient. Following Sollitt and Cross (1972) and Madsen (1974), / is evaluated from the following equation,
Jo Jv ' ^ l^l 2 + ^ l"il3 ^ f=^"r — "
dtdV
1
w
Jo
>
30
JvnelutfdtdV
for a porous structure of volume V under a wave-cycle of period T. Please note that Uj is taken to be the real part of the seepage velocity and therefore, an interative procedure is needed to evaluate / . 4.2. Solutions
based on eigenfunction
expansions
Based on this irrotational and linear approximation, Sollitt and Cross (1972) presented a model to analyze wave interaction with vertically sided porous structures. This work was later extended by Dalrymple et al. (1991) to include oblique incident waves. Dalrymple et al. (1991) considers the interaction of a gravity wave train with a single homogeneous, isotropic, porous structure of width b between two semi-infinite fluid regions of constant depth h. The wave
176
/. J. Losada
field outside the structure can be specified by velocity potentials $ ! in the seaward region and $3 in the leeward region of the breakwater by specifying the well-known linear boundary-value problem for water waves in constant water depth. In the rigid porous medium, region 2, a boundary value problem can be defined using Eqs. (24), (28), and (29) and adding the kinematic bottom boundary condition 8<&2l'dz = 0 in z = —h. Each of the boundary value problems is formulated in terms of linear homogeneous equations. Separation of variables leads to Sturm-Liouville problems where the potentials may be expressed in terms of an eigenfunction expansion. Since the solution in adjacent regions must be continuous at each interface, continuity of mass flux and pressure at x = 0 (interface between regions 1 and 2) and at x = b (interface between regions 2 and 3) is required. These conditions may be expressed as: $ix = ne$2x ,
$1 = (s - i / ) $ 2 at x = 0 ,
(31a)
®ix = ne$2x ,
$1 = (s - i / ) $ 2 at x = b,
(31b)
and
where the continuity of pressure is derived from the Bernoulli equation Eq. (25). Substituting the potentials into the matching conditions in Eqs. (31a) and (31b), a system of equations is obtained. Unknowns are the complex amplitudes of the progressive and evanescent modes in the potentials $1, $2, and $3. Applying the orthogonality of the eigenfunctions over the water depth in the 3 regions results in a simpler system. Once the system of equations is solved together with Eq. (30) and the corresponding dispersion relationships, the potential and therefore, the flow is completely defined. Dalrymple et al. (1991) presents the variation of reflection and transmission coefficients for several rectangular geometries, finite, semi-infinite and infinite breakwater with an impermeable wall considering several relative water depths and analyzing the influence of wave incidence. They found that a minimum reflection coefficient occurs for different angles of incidence depending on the / value. It has to be pointed out that results in this work have been obtained assuming a given constant / . The eigenfunction approach has also been applied to crowned breakwaters (Losada et al, 1993) or to submerged porous steps (Losada et al., 1997a). The complex nature of the dispersion equation as Eq. (29) leads to two particular difficulties when Sollitt and Cross (1972) model is used in conjunction
Recent Advances in the Modeling of Wave
177
with eigenfunction expansions technique. First, it is difficult to locate the complex roots of the dispersion relation by standard numerical methods. Second, the vertical eigenfunction problem is not self-adjoint and standard expansions theorems do not apply. These problems have been discussed by several authors (f.i. Dalrymple et al., 1991). Mclver (1998) presents a method that allows the explicit calculation of the roots of the complex dispersion relation and uses the theory of non-self-adjoint differential operators to show how the formal construction of eigenfunction expansions can be carried out for the interaction of water waves with porous structures. In order to consider different geometries generalizing the theory of Sollitt and Cross (1972), Sulisz (1985) developed a boundary element method to investigate wave transmission and reflection from a multilayered breakwater with arbitrary shape. Based on a linearized long wave theory, Massel and Mei (1977) and Massel and Butowski (1980) were the first to consider random wave interaction with permeable structures. Following Dalrymple et al. (1991), Losada et al. (1997b) considered the interaction of directional random waves with vertical permeable structures. Using an eigenfunction expansion, Losada et al. (1997b) simulated the transformation of a given incident spectrum in the vicinity of the partially reflecting structures. The influence on the results of the structure's geometry, permeable material characteristics and incident wave spectrum is analyzed. Further information regarding linear solutions for monochromatic waves based on a linearized version of Sollitt and Cross (1972) equations can be found in Chwang and Chan (1998). 4.3. Mild-slope
equation
Within the framework of linear wave theory, Berkhoff (1972) proposed a twodimensional theory which can deal with large regions of refraction and diffraction. This new equation, the mild-slope equation, has been extensively used for wave propagation modeling. In order to consider wave propagation on porous slopes and wave interaction with trapezoidal permeable submerged breakwaters, Rojanakamthorn et al. (1989), Losada et al. (1996a) and Mendez et al. (2000) present extended versions of the mild-slope equation. The extension of the mild-slope equation for permeable layers is derived by multiplying the Laplace equation by its correspondent vertical eigenfunctions,
178
/. J. Losada
Fig. 2. Schematic description of the submerged permeable structure geometry.
MQ{Z) and Po(z), expressed in terms of the propagating mode only, neglecting evanescent modes and integrating over depth. Following Losada et al. (1996a), the new governing equation is: /
M0(z) (v2h§2 +
J-h+ah
\
dz2
dz + ne(s — if)
r-h+ah
p 0 (z)(V£$ 4 + ^ | d z = 0,
(32)
where $ 2 = (p(x,y)M0(z)
and
$4 =
ip(x,y)PQ{z),
(33)
and $2 is the velocity potential in the fluid region above the permeable layer or submerged breakwater, $4 the potential inside the permeable region, y> is the complex amplitude of the water surface and V/, = (d/dx, d/dy). The boundary and matching conditions for variable depth are: • Combined free surface boundary condition, r-ll dz
$ 2 = 0 at 2 = 0.
(34)
g
• Bottom boundary condition, <9$4 dz
Vh • W - $ 4 = 0 at z =
-h.
(35)
Continuity of mass flux, <9$4 oz
+ Vh-hVh-$>4
at z = - / i + a/i.
(36)
Recent Advances in the Modeling of Wave
179
• Continuity of pressure, $2
(37)
(s - i/)$4 at z = -h + ah .
Integrating Eq. (32) using the boundary and matching conditions and finally the mild-slope assumption, the following equation is obtained, V h • (XVfc •
(38)
where X
(£) [£
M$(z)dz +
-h-\-a.h
ne(s-if)
r
h-\-cth
P*(z)dz
(39)
J-h and where the term iu>fifD
SD =
where Cg is the group velocity, TR is the real part of To, tan£ is the equivalent bottom slope at the breaking point which is defined as a mean slope in the distance 5heftb offshore the breaking point and, v
hef
vs =0.4(0.57 + 5 . 3 tanC),
(41)
vr = QA-,M
h.ef
where the subscript b means the value at the breaking point. hef is the effective water depth over the porous layer as defined in Losada et al. (1997a). To solve the problem, a finite difference scheme and proper boundary conditions at the domain boundaries are used. The resulting system of equations
180
/. J. Losada
including Eq. (30) can be simultaneously solved. Assuming a breaking criteria, fo is calculated using the expression in Rojanakamthorn et al. (1990). In order to arrive to a solution, an iterative procedure has to be carried out. Losada et al. (1996b) extends the model to consider the interaction of directional random waves with submerged breakwaters. The influence of structure geometry, porous material properties and wave characteristics including oblique incidence, on the kinematics and dynamics over and inside the breakwater is considered in Losada et al. (1996a, 1996b) and Mendez et al. (2000). Models are validated against experimental data. Figure 3 shows root-mean-square wave height Hrms evolution in a submerged permeable breakwater. The numerical model results obtained in Mendez et al. (2000) are compared with experimental data (Rivero et al., 1998). The submerged breakwater with 1:1.5 slopes on both sides has a crown width of 0.61 m is constructed on an impermeable core and an armor layer of quarrystones with a mean weight of 25 kg. The structure is placed on a 1:15 rigid slope bottom. The incident irregular wave characteristics are given by 0.396 m and Tp = 4 s. Applying Eq. (11) and with ne = 0.4, the rest rmsi of the permeable material characteristics are: Kp = 2.5 x 10~ 5 m 2 , Cj = 0.3, s = 1. The resulting linearized friction coefficient is / = 2.49. Results show very good agreement even under breaking conditions. In front of the structure, wave reflection induces a modulation of the wave height. Wave '
1
'
1
'
1
'
i——^*
H.„.,= 0.396 m T„=4.0s _
rms.1
1
•
1
/•
*^^\
o.io
1
p
•\
'
1
1
1
'
1
'
_ -
-
a =1.5 |R 1=0.02 * = 4 5 ° - K^ 2.5E-05 m2 Cf= 0.3 _ f=2.49 s=l
Fig. 3. Wave transformation by a permeable submerged breakwater. Comparison of experimental and numerical results.
Recent Advances in the Modeling of Wave
181
breaking takes place on the crest. A modulation of the wave height is also visible leewards the submerged breakwater due to the reflection induced by the bottom slope. Mase and Takeba (1992) extends the mild slope equation, deriving timedependent and time-independent wave equations for waves propagating over porous rippled beds. By using the time-independent equation, the Bragg scattering is examined in a one-dimensional case showing that energy dissipation in the porous layer contributes to smaller reflected and transmitted coefficients than those in the case of an impermeable rigid rippled bed. 5. Shallow Water Models 5.1.
Introduction
There are numerous situations where accurate computations of the wave field on permeable beaches or around permeable structures are not possible using the mathematical models presented in previous sections. Nonlinearity is an important feature in the process on wave interaction with most coastal structures usually located between intermediate and shallow water depths. For example, in order to model the generation of higher harmonics on regions of abrupt depth variations such as crowns of permeable submerged structures (Losada et al., 1997a), the mild-slope equation is limited by the use of linear theory. A review on shallow water models can be found in van Gent (1995). 5.2. Recent
developments
Using a perturbation method, Cruz et al. (1992) derived a set of time-dependent nonlinear equations for one-dimensional wave transformation on porous beds. Since these models include the leading order of nonlinearity, they are able to generate higher harmonics on the shallow water region. However, the inherent dispersivity is weak and consequently the frequency-dependent wave decomposition beyond submerged breakwaters cannot be reproduced. Following the approach by Madsen et al. (1991) for impermeable beds, Cruz et al. (1997) derived a set of Boussinesq equations over a porous bed of arbitrary thickness with an underlying solid bottom of arbitrary depth. After determining the governing equations and boundary conditions for the three-dimensional wave motion, assuming incompressible and irrotational flow in both the fluid and permeable layer, a set of time-dependent, vertically-integrated equations is
182
/. J. Losada
derived containing the leading order of nonlinearity. The weak dispersivity of Boussinesq-type equations is corrected by adding dispersion terms to the basic momentum equations and matching the resulting dispersion relation with that of appropriate theory. The equations of motion inside the porous layer include resistance terms following the approach by Sollitt and Cross (1972). The permeable material parameters were extrapolated from values tabulated in Sollitt and Cross (1972). Numerical results are obtained for wave propagation on a horizontal bottom of uniform thickness with good agreement. Liu and Wen (1997) derived a two-dimensional fully nonlinear, diffusive and weakly dispersive set of equations for long wave propagation in a shallow porous medium. Using analytical perturbation solutions as well as numerical solutions, the one-dimensional equations are used to study the tide-induced free surface fluctuations in a porous region and the transmission and reflection of solitary waves by a rectangular porous breakwater. The last is calculated considering the Boussinesq approximation carrying out a linearization process inside the porous breakwater to convert the nonlinear resistance formula to a Darcy-type resistance. Numerical results are compared with experimental data (Vidal et al, 1988) observing an excellent agreement. 5.3. Diffraction
by porous
structures
Diffraction of waves by a solid breakwater has received a considerable amount of attention. However, the role of breakwater porosity on the wave diffraction process has not been addressed until recently. Based on the linear potential wave theory, Yu (1995) developed a porous breakwater diffraction model. This model was extended to waves of arbitrary incidence (Yu and Togashi, 1996; Mclver, 1999), but requires that the breakwater be thin compared to the incident wave length. Lynett et al. (2000) presents a model based on depthintegrated equations suitable for weakly nonlinear, weakly dispersive transient waves propagating in both variable-depth open water and porous region. In this first work, the model is applied to analyze solitary wave interaction with vertically-walled porous structures in a horizontal bottom. In the open water region, the model employs the generalized Boussinesq equations presented originally by Wu (1981). The equations are expressed in terms of the free surface displacement ( and the depth-averaged velocity potential <j>. In dimensional form, the equations are given as:
Recent Advances in the Modeling of Wave
^ | + V.[(C + W ] = 0 , ^ + 2 ( W
+
K - ^ V . (
W )
+
183
(42) - - V ^ = 0,
(43)
where /i is the local water depth, g is gravity and V = (d/dx,d/dy) the horizontal gradient. Using 6, the depth-averaged velocity u can be calculated by: u = V<j>. (44) These equations are valid only for weakly nonlinear and dispersive waves, limiting the application to waves satisfying 0{A/h) = 0(kh)
A _ _ V [ ( C + / l ) - v c ] - y ^ v 2 C = o, h?
(45) (46)
Velocity in the porous region is given as: u = -KViP .
(47)
The resolution of the problem requires reflective and radiation conditions at the exterior boundaries and matching conditions along the interface between the fluid and porous region. The matching conditions are given as:
C|+=CI-, n-V<|+=u-VC|-,
u\+=u\-,
V-(n-u)|+ = V-(n-il)|_,
(48) (49)
where the sign denotes opposite sides of the interface and n is the unit normal vector. Velocities are evaluated using Eqs. (46) and (47).
184
/. J. Losada
Using a high-order finite difference scheme, the numerical model is used for both ID and 2D problems. Considering solitary wave interaction with a permeable vertical structure considerably simplifies our understanding of the wave diffraction-transmission process. To validate the model and compare the different mechanisms, a comprehensive set of experiments was performed in a wave tank with a porous and impermeable breakwater perpendicular to solitary wave incidence. Water depth, wave height and gravel diameter were varied and free surface was recorded in a dense grid covering the region in front of the structure and in the shadow zone. Results show that the model predicts wave height, wave form and arrival time excellently. Furthermore, the model and experimental results are useful for evaluating the differences between the wave field in the shadow zone comparing the porous breakwater case with the solid breakwater case. Figure 4 presents the results of the numerical simulation of the interaction of a solitary wave passing a solid (left) and a porous (right) detached breakwater. The snapshots correspond to spatial profiles of a normally incident solitary wave with A/h = 0.1 interacting with a breakwater, with a length of 5 water depths and a width of 80 water depths. For the porous breakwater, the scaled rock size has the following characteristics Dso/h = 0.2, ne = 0.5 m ap = 1100, and (3p = 0.81. The first two snapshots (a) and (b) show the solitary wave approaching the detached breakwater. In Figs. 4(c) and 4(d), the wave height at the front face is at a maximum, while Figs. 4(e) and 4(f) show the reflected waves at the beginning of diffraction behind the breakwater. The analysis of the shadow zone in Figs. 4(e), 4(f), 4(g), and 4(h) shows that less energy is diffracting to form a wave with a circular crest line in the porous breakwater case than in the solid breakwater case. The main difference between both cases is that most of the energy that diffracts in the solid breakwater case to form this circular wave, diffracts into the transmitted wave front in the porous breakwater case. This is due to the fact that in the shadow zone of the porous breakwater, wave diffraction occurs in two ways. A wave with circular crest lines is created since part of the wave energy is diffracting into the calm water behind the transmitted wave in the same form as diffraction occurs behind a solid breakwater. The second part of diffraction takes place because wave energy diffracts into the transmitted wave front from the incident wave front due to the discontinuity of wave amplitude. The relative importance of each of the mechanisms will depend on the incident wave characteristics, breakwater geometry and permeable material characteristics.
Recent Advances in the Modeling of Wave
185
(b)
(a) n
m
JJ )
(d)
(c)
•
1
in (f)
m
Ijj
Fig. 4. Numerical simulation of solitary wave interaction with detached impermeable and permeable breakwaters.
186
/. J. Losada
6. Short Wave-Averaged Flow 6.1.
Introduction
6.2. Time and depth-averaged
equations
The continuity and Navier-Stokes equations describing the fluid flow may be written in terras of the instantaneous vertical velocity component w* and the viscous stress tensor r,*.- as: du* dxi du* dt
d(u*u*) dxi
dw* _ dz
d(u*w*) dz
di-P^ij+rti} dxi
dw* , d(w*u*) , d(w*2)] dt dxi dz
d[-p* + pgz] dz
^ dr*jz dz dr* dxi
J = 1,2
(51)
dr*zz dz
where the viscous shear stress tensor can be expressed in terms of the instantaneous velocity field,
Inside the porous medium, the velocity and pressure fields can be decomposed according to Eq. (14). Each of these components can be resolved in a mean component (depth and time-averaged value) and a deviation from the mean, Uj
=
Ui+Ui,
= Ui + ui, A = Ut + u\,
«i
(54)
where Ui is the seepage current, \n is the seepage oscillatory flow, U? is the spatial fluctuation of the current, U\ is the temporal fluctuation of the current, u\ is the spatial fluctuation of the oscillatory flow and u\ is the temporal fluctuation of the oscillatory flow. In order to derive the time and depth-averaged equations, Losada (1996) carried out the following operations, • The continuity equation, Eq. (50), is averaged over a finite volume of porous medium and in a time scale smaller than the characteristic wave period resulting in a continuity equation in terms of the seepage components.
Recent Advances in the Modeling of Wave 187
• This equation is integrated from the bottom z = — h to the phreatic surface z = j). Applying Leibniz's rule using the kinematic boundary conditions and averaging over a wave period yields,
on
dim±m
dt
=
dxi
K
'
'
where
<56)
"'-TC {ah £«*)"•
Following a similar procedure with the momentum equations (see Losada, 1996 for details) and assuming that no correlation exists between the seepage flow and the fluctuations and that the correlation between the mean and oscillatory flow is also very weak, the following equations can be obtained, d[pUj{fj + h)}
— m —
dfj =
d
~"{T]+h)dVi-d^i d
dx,
{Sij+5 +5
t
« «>
p(fj + h)(U!U° + U*U*) +TJ+Rj,
(57)
where
P
1 _
S^ = / (pSij + puiUj) dz- -2' pg(rj + h)2Sij , J-h S^ = /
pufuj dz ,
(58) (59)
J —h
%,= T pS$dz,
(60)
J —h
Sij is the (i,j) component of the stress tensor representing the excess of momentum flux due to seepage magnitudes. S*j are S\j the stress tensor components due to the spatial and temporal fluctuations. These two terms may be considered lateral mixing mechanisms induced by the temporal and spatial fluctuations due to pore irregularities. The horizontal force term due to the oscillatory motion T; is given as: Ti = (pU-pg(n + h))^-Pg(h
+ n)§^.
(61)
188
/. J. Losada
The frictional terms on the free surface F(x,y,z) z) = 0 have the following expression, R
o = -^-.fnjdz
= 0 and bottom
+ Vf\VF\+^\WB\.
B(x,y,
(62)
Equations (55) and (57) may be simplified under certain circumstances to obtain other equations that are important for practical applications. Assuming horizontal bottom, neglecting viscous forces (Rj = 0), and using Eq. (55), the momentum equations can be expressed as: f
dfj\
d2fj
1
dxi \
dxi)
dt2
p dxidxj
S^Sij
+ S^+Slj
d
d2
+ lUfUj + UUJl}.
(64)
Equation (64) governs the forced long wave motion in a porous medium. 6.3.
Applications
The application of the time and depth-averaged equations requires expressions for the wave-averaged quantities, mass flux, radiation stress, etc. in terms of wave height, wave period and water depths. Assuming irrotational flow and based on the potential <j) derived from Sollitt and Cross (1972) theory, expressions correct to the second order have been derived for porous media flow in terms of the seepage velocity. Mendez et al. (2000) provides general expressions of mass transport, radiation stress and energy flux for wave propagation inside or above a porous layer in terms of some shape functions defined accordingly. Using an approximation, Mendez et al. (2000) applied a simplified ID version of Eq. (57) considering the forcing of the radiation stress of the seepage component in order to analyze the mean water level variations induced by the presence of a submerged permeable breakwater. Figure 5 presents numerical and experimental results of wave height and mean water level variation for a submerged permeable breakwater. A description of the experimental conditions can be found in Rivero et al. (1998). Results show a modulation of the mean water level in front of the structure imposed by reflection. A set-down can be clearly observed at the front slope or above the breakwater crest just before the maximum wave height is reached before breaking. After breaking, a slightly modulated set-up is clearly observed leewards the structure in both the experimental and numerical results. Results
Recent Advances in the Modeling of Wave
189
0.80
0.60
•J»«L»
^ * ^ = * ^ * 0.20 Hr= 0.435 m T = 4.0 s r\ = -1.5 cm 0.00
L
4.00
V
2
2.5E-05 m
f 0.3
C
S=l
0.0, f=1.30
2.00 0.00
- - • --- - - •
|R| 0.07 V
0°, f=1.24
mi 0.07 V -60° . f=1.34
9.0
x(m) Fig. 5. Wave height and mean water level variation in a submerged breakwater under breaking conditions.
point out that the set-up induced by a permeable submerged breakwater is due to radiation stress gradients induced by dissipation associated with both breaking and friction. In general, the set-up due to breaking is much more important than the one produced by the porous material. However, the relative magnitude of the different contributions requires further exploration due to the lack of roller effects, flow separation or turbulent effects in this first approach to the problem. Baquerizo and Losada (1998) used Eqs. (55) and (57) to analyze how dissipation inside an infinitely long vertical breakwater in a constant water depth generates an along-breakwater current inside the structure. For an obliquely incident wave train, dissipation of energy inside the pores produces a radiation stress gradient which is balanced by the frictional and the diffusive terms in
190
/. J. Losada
the equation. The current driven inside the porous medium is transferred by turbulent diffusion seawards and leewards the structure. The solution of the equations requires an expression for the radiation stresses associated with the fluctuations. Baquerizo and Losada (1998) based on a Boussinesq approach give approximated expressions introducing eddy viscosity coefficients which take into account the effect, over the mean velocity, of the turbulence induced by the spatial and temporal fluctuations respectively. Approximate expressions for these terms need to be explored further. Losada et al. (1998) analyzed the mean flows in vertical rubble-mound structures. Using an analytical approach based on Sollit and Cross (1972) theory, it is shown that waves impinging on rubble mound breakwaters and seawalls induce a mean flow within the breakwater analogous to the so-called undertow within the surf zone. The mean flow is expressed in terms of a mean stream function which is analytically derived and is based on the mass flux balance between the incident, reflected and transmitted waves. It can be shown that the rapid decrease in the Eulerian mass transport results in a mean vertical velocity that capable of inducing a mean flow that exits mainly at the toe of the structure. The higher the reduction in mass flux, the stronger the mean return current. *F in a vertical rubble-mound breakwater |m|=0.20
_
z/h
Mean horizontal velocity profile
-m V
n
:*#* Return flow
If •*••••
M3
-|m|=0.50 -]m|=p.20 -•|iii=0.U,-
i
Fig. 6. Streamlines and mean horizontal velocity profiles in a finite breakwater, b/h = 1, kh = 0.34, and ne = 0.4.
Recent Advances in the Modeling of Wave
191
The upper panel of Fig. 6 shows the calculated streamline patterns for a finite porous breakwater. The admittance m = neT/(s — if) where V is the complex wave number in the porous medium takes into account the hydraulic characteristics of the structure. Increasing friction results in diminishing admittance. The lower panel presents a qualitative plotting of the horizontal mean velocity profiles for different admittances. The results indicate that for \m\ = 0.5, the smallest friction considered, the influence of the presence of the structure on the mean velocities is almost negligible. However, decreasing admittance turns out in higher return flow due to increasing mass flux decay in z =0. For 177i| = 0.14, the velocity profile presents a large curvature and showing a net flux in the direction of wave propagation near the top of the structure. 7. Modeling Based on the Navier- Stokes Equations In order to overcome most of the limitations associated with previous models, important efforts have been made in the last few years to develop a tool capable of successfully simulating the free surface of breaking waves on permeable structures. Three succesful examples are the models SKYLLA (van Gent, 1995), VOFbreak (Troch and de Rouck, 1998) and COBRAS (Liu et al, 1999, 2000), all based on 2D Navier-Stokes equations for the fluid and porous regions and making use of the VOF method to track the free surface. COBRAS (COrnell BReAking wave and Structure) has been initially developed to track the free surface movement and to describe the turbulence generated by the wave breaking process on slopes (Lin and Liu, 1998a, 1998b). The breaking waves numerical model is based on the Reynolds Averaged Navier-Stokes (RANS) equations. Although the model has only been applied to two-dimensional problems, the complete three-dimensional formulation will be given herein. For a turbulent flow in the fluid region, the velocity field and pressure field can be divided into two parts: The mean (ensemble average) velocity and pressure and (u$) and (p), the turbulent velocity and pressure u\ and p'. Thus, Ui = (ui) +u'i,
p= (p)+p'
(65)
in which i = 1,2,3 for a three-dimensional flow. If the fluid is assumed incompressible, the mean flow field is governed by the Reynolds Averaged
192
/. J. Losada
Navier-Stokes equations, 9(uj) dxi
dt
(uj)
%) = _19(p) + dxj p dxi
(66)
0,
{ l
ld{rtj) p dxj
djuty) dxj
(67)
in which p is the density of the fluid, gi the zth component of the gravitational acceleration, and mean molecular stress tensor (r^) = 2p{<7ij) with p being the molecular viscosity and (cr^) = \{ ^ " ' ' + QJ ) the rate of strain tensor of the mean flow. In the momentum equation as in Eq. (67), the influence of the turbulent fluctuations on the mean flow field is represented by the Reynolds stress tensor /o(w^u'). Many second-order turbulence closure models have been developed for different applications which have been summarized in a recent review article (Jaw and Chen 1998a, 1998b). In the present model, the Reynolds stress p(u^u') is expressed by a nonlinear algebraic stress model (Shih et al., 1996; Lin and Liu (1998a, 1998b):
p(uiuj)
= ^pkSij
~P~2
+ C2
+ C,
It2 (d(m) -Capdxi
Ci
d(ui) d(ui) dxi dxj
9{UJ)
dxi d{uj) d(ui) dxi dxi
2 d(ui) d(uk) 3 dxk
dxi
d(uj) d{uj) _ 1 d{ut) d(ui) dxk
dxk
3 dxf. dxk
fd(uk)
d{uk)
V dxi
dxj
1 d(ui) djm) 3 dxk dxf.
5 lJ
(68)
in which Cd, C\, C2, and C3 are empirical coefficients, Sij the Kronecker delta, k = ^{u'iii'i) the turbulent kinetic energy, ande v((d^L)2) ^ n e dissipation rate of turbulent kinetic energy, where v = p/ p is the molecular kinematic viscosity. It is noted that for the conventional eddy viscosity model C\ = C2 = C3 = 0 in Eq. (68) and the eddy viscosity is then expressed as« ( = Cd — - Compared with the conventional eddy viscosity model, the nonlinear Reynolds stress model in Eq. (68) can be applied to general anisotropic turbulent flows.
Recent Advances in the Modeling of Wave
193
The governing equations for k and s are modeled as (Rodi, 1980; Lin and Liu, 1998a, 1998b): dk dt de
m+
. . dk dxj
d dxj
vt i\crk + v
de
_d_ dxj
vt_
^dx-
dk dxj
d(uj) «
de
-a 2e
;
•
>
•Cu\rH
dx-j
djuj) dxj
«
(69)
dxj
d(ui) dxj
8(UJ) dxi
(70)
k
in which o^, a£, C\e, and Cie are empirical coefficients. In the transport equation for the turbulent kinetic energy, Eq. (69), the left-hand side term denotes the convection while the first term on the right-hand side represents the diffusion. The second and the third term on the right-hand side of Eq. (69) are the production and the dissipation of turbulent kinetic energy respectively. The coefficients in Eqs. (68) to (70) have been determined by performing many simple experiments and enforcing the physical realizability; the recommended values for these coefficients are (Rodi, 1980; Lin and Liu, 1998a, 1998b): Cd =
1 3 V 7.4 + 5 m a x
C2 = -
1 58.5 + D ^ '
C l £ = 1.44,
C 2e = 1.92,
Ci
C,
1 185.2+ D2 (71) 370.4 + D ^ x '
<7fc = 1.0,
<7e = 1.3,
where 5,max — 7 max | 1^1 (repeated indices not summed) and Dn - maxf! Appropriate boundary conditions need to be specified. For the mean flow field, the no-slip boundary condition is imposed on the solid boundary and the zero-stress condition is required on the mean free surface by neglecting the effect of airflow. For the turbulent field near the solid boundary, the log-law distribution of mean tangential velocity in the turbulent boundary layer is applied so that the values of k and e can be expressed as functions of distance from the boundary and the mean tangential velocity outside the viscous sublayer. On the free surface, the zero-gradient boundary conditions are imposed for both k and e , i.e., §^ = | p = 0. A low level of k for the
194
/. J. Losada
initial and inflow boundary conditions is assumed. The justification for this approximation can be found in Lin and Liu (1998a, 1998b). In the numerical model, the RANS equations are solved by the finite difference two-step projection method (Chorin, 1968). The forward time difference method is used to discretize the time derivative. The convection terms are discretized by the combination of central difference method and upwind method. The central difference method is employed to discretize the pressure gradient terms and stress gradient terms. The VOF method is used to track the free surface (Hirt and Nichols, 1981). The transport equations for k and e are solved with the similar method used in solving the momentum equations. Detailed information can be found in Kothe et al. (1991), Liu and Lin (1997), and Lin and Liu (1998a, 1998b). The mathematical model described above has been verified by comparing numerical results with either experimental data or analytical solutions. The detailed descriptions of the numerical results and their comparison with experimental data can be found in Liu and Lin (1997) and Lin and Liu (1998a, 1998b). The overall agreement between numerical solutions and experimental data was very good. The flow in the porous domain is described using Eqs. (21), (22), and (11). In order to couple the flow inside and outside the permeable structure, Liu et al. (1999) applies continuity of the mean and averaged velocity and pressure across the interface of porous media and outside flow. Strictly speaking, the outside mean (ensemble-averaged) flow is not equivalent to the spatiallyaveraged flow in porous media since the latter may still contain turbulent fluctuations. However, as previously explained, these turbulent fluctuations are in general negligible. Please note that the turbulence model is not solved in porous media. Therefore, the evaluation of the turbulence kinetic energy needs a special treatment. For details, see Liu et al. (1999). The model is suitable to analyze wave interaction with emerged or submerged permeable or solid structures, both for breaking or nonbreaking conditions. Figure 7(b) shows the comparison of calculated and experimental time histories of free surface displacement before and after the wave passes a trapezoidal porous structure, Fig. 7(a). Reasonably good agreements are obtained. In Fig. 7(c), the vertical and horizontal components of the velocity at two locations are shown. Both experimental and numerical results show that because
Recent Advances in the Modeling of Wave
195
0.6 0.5 0.4 measurement sections
It 0.3 0.2 0.1
-3
-2.5
-2
-1.5
-1
-0.5 x(m)
0
0.5
1
1.5
(a)
free surface displacement at x=0.24 m (e), 0.33 m (f), 0.42 m (g), and 1.42 m (h) E u
10 0
-10 20 E o
10 0
(f)
/
-10 20 E o
10 0
(g)
\
i\
A
-10 20 E 10
(b) Fig. 7. (a) Wave transformation above a trapezoidal submerged breakwater. Location of the free surface gauges, (b) Free surface displacement at different locations. Comparison between experimental and numerical results, (c) Horizontal and vertical velocity time histories at x = 0.24 and at different depths. Experimental and numerical results.
196
/. J. Losada u & v at x=0.24 m and y=0.30 m (a), 0.16 m (b), 0.12 m (c), 0.10 m (d) and 0.08 m (e)
-0.5' 4
' 6
' 8
=—' 10 t(m)
3
12
' 14
-=>
(c)
Fig. 7.
(Continued)
of the presence of the structure, the horizontal velocity near the structure is enhanced while the vertical velocity is reduced. The agreement is again quite good. 8. Conclusions Very important progress has been achieved in the field of modeling wave and permeable structure interaction in the last years. The modeling is based on the coupling of two models describing the flow acting on the structure and through the porous structures. Several new equations including the resistance forces in the porous medium have been derived covering an ample range of applications. The accuracy of the modeling of wave and permeable structure interaction relies greatly on some constants depending on the flow. Even if some progress has been carried out to determine predictive formulas for these constants under oscillatory flow conditions, uncertainty is still present especially if these models are considered to be an alternative to physical modeling in the design of coastal structures. Equations for wave and structure interaction have evolved in parallel with equations for wave-propagation in fluids. Modified Boussinesq equations or other kind of shallow water equations are currently available. However, the
Recent Advances in the Modeling of Wave
197
application by the engineering community is limited. From the engineering point of view, there is still additional work to do in order to include the effect of permeable structures in the wave propagation modeling. For wave and permeable structure interaction where the numerical domains are relatively small, models based on the RANS equations seem to be the most suitable. However, this kind of model requires further work and validation before being useful for engineering applications. 9. Future Work Further research on the determination of predictive expressions for the porous flow parameters under oscillatory flow conditions is needed. It is clear that in the near future, models solving the Navier-Stokes equations have an enormous potential to analyze wave and permeable structure interaction. Therefore, COBRAS should be improved in several aspects. The enforcement of the continuity of the mean and averaged velocity and pressure across the interface results in unrealistic velocity information just outside the porous media. Just at the porous media surface, the flow forms turbulent jets and wakes that in most practical applications will be quickly mixed within a short distance. If detailed velocity information at locations very close to the porous surface is needed, the present approximation results should be improved. This could be the case for the development of structure stability models where the forces on individual units have to be determined. Therefore, the modeling of turbulence inside the structure should be explored. Furthermore, different turbulence closure models need to be considered. The extension of COBRAS to three-dimensions would imply a considerable improvement to the current model. Acknowledgment The financial support from the Spanish Comision Interministerial de Ciencia y Tecnologia (CICYT) through grant MAR99-0653 is gratefully acknowledged. References Baquerizo, A. and M. A. Losada (1998). Longitudinal current induced by oblique waves along coastal structures. Coastal Eng. 35: 211-230. Bear, J. (1972). Dynamics of Fluids in Porous Media. American Elsevier, New York. Berkhoff. J. C. W. (1972). Computation of combined refraction-diffraction. Proc. 13th International Conference on Coastal Eng., Vancouver, 471-490.
198
/. J. Losada
Burcharth, H. F. and O. H. Andersen (1995). On the one-dimensional steady and unsteady porous flow equations. Coastal Eng. 24: 233-257. Carman, P. C. (1937). Fluid flow through granular beds. Trans. Inst. Chem. Eng. 15: 150-166. Chorin, A. J. (1968). Numerical solution of the Navier-Stokes equations. Math. Comput. 22: 745-762. Chwang, A. T. and A. T. Chan (1998). Interaction between porous media and wave motion. Ann. Rev. Fluid Mech. 30: 53-84. Cruz, E., M. Isobe and A. Watanabe (1997). Boussinesq equations for wave transformation on porous beds. Coastal Eng. 30: 125-154. Dalrymple, R. A., M. A. Losada and P. Martin (1991). Reflection and transmission from porous structures under oblique wave attack. J. Fluid Mech. 224: 625-644. Ergun, S. (1952). Fluid flow through packed columns. Chem. Eng. Progress 48(2): 89-94. van Gent, M. R. A. (1993). Stationary and oscillatory flow through coarse porous media. Communications on Hydraulic and Geotechnical Engineering. TU Delft. ISSN 0169-6548. Report 93-9. van Gent, M. R. A. (1991). Formulae to describe porous flow. Communications on Hydraulic and Geotechnical Engineering. TU Delft. ISSN 0169-6548. Report 92-2. van Gent, M. R. A. (1995). Porous flow through rubble-mound material. J. Wtrway., Port, Coast, and Oc. Eng. ASCE. 121: 181. Gu, Z. and H. Wang (1991). Gravity waves over porous bottoms. Coastal Eng. 15: 695-524. Hall, K. R., G. M. Smith, and D. J. Turcke (1995). Comparison of oscillatory flow and stationary flow through porous media. Coastal Eng. 24: 217-232. Hirt, C. W. and B. D. Nichols (1981). Volume of Fluid (VOF) method for the dynamics of free boundaries. J. Comp. Phys. 39: 201-225. Kozeny, J. (1927). Uber kapillare Leitung des Wassers im Boden, Sitzungber. Akad. Wiss. 136: 271-376. Jaw, S. Y. and C. J. Chen (1998a). Present status of second-order closure turbulence model. I: overview. J. Eng. Mech. 124: 485-501. Jaw, S. Y. and C. J. Chen (1998b). Present status of second-order closure turbulence models. II: application. J. Eng. Mech. 124: 502-512. Kothe, D. B., R. C. Mjolsness, and M. D. Torrey (1991). RIPPLE: A Computer Program for Incompressible Flows with Free Surfaces. Los Alamos National Laboratory, LA-12007-MS. Lin, P. and P. L.-F. Liu (1998a). A numerical study of breaking waves in the surf zone. J. Fluid Mech. 359: 239-264. Lin, P. and P. L.-F. Liu (1998b). Turbulence transport, vorticity dynamics, and solute mixing under plunging breaking waves in surf zone. J. Geophys. Res. 103: 15677-15694.
Recent Advances in the Modeling of Wave
199
Liu, P. L.-F. and P. Lin (1997). A Numerical Model for Breaking Wave: The Volume of Fluid Method. Research Report No. CACR-97-02. Center for Applied Coastal Research, Ocean Eng. Lab., University of Delaware, Newark, DE. Liu, P. L.-F. and J. Wen (1997). Nonlinear diffusive surface waves in porous media. J. Fluid Mech. 347: 119-139. Liu, P. L.-F., P. Lin, K.-A. Chang, and T. Sakakiyama (1999). Numerical modeling of wave interaction with porous structures. J. Wtrway., Port, Coast, and Oc. Eng. ASCE 125: 322-329. Liu, P. L.-F., T.-J. Hsu, P.-Z, Lin, I. J. Losada, C. Vidal, and T. Sakakiyama (2000). The Cornell breaking wave and structure (COBRAS) model, in: Proc. Coastal Structures 99, ed. I. J. Losada. 169-175. Losada, I. J. (1996), Wave Interaction with Permeable Structures. Ph.D. Thesis. Department of Civil Engineering. University of Delaware. 230 pp. Losada, I. J., R. A. Dalrymple, and M. A. Losada (1993). Water waves on crown breakwaters. J. Wtrway., Port, Coast, and Oc. Eng. ASCE 119(4): 367-380. Losada, I. J., M. A., Losada, and F. L. Martin (1995). Experimental study of waveinduced flow in a porous structure. Coastal Eng. 26(1-2): 77-98. Losada, I. J., R. Silva, and M. A. Losada (1996a). 3D nonbreaking regular wave interaction with submerged breakwaters. Coastal Eng. 28(1-4): 229-248. Losada, I. J., R. Silva, and M. A. Losada (1996b). Interaction of nonbreaking directional random waves with submerged breakwaters. Coastal Eng. 28(1-4): 248-265. Losada, I. J., M. D. Patterson, and M. A. Losada (1997a). Harmonic generation past a submerged porous step. Coastal Eng. 3 1 : 281-304. Losada, I. J., R. Silva, and M. A. Losada (1997b). Effects of reflective vertical structures permeability on random wave kinematics. J. Wtrway., Port, Coast., and Oc. Eng. ASCE 123(6): 347-353. Losada, I. J., R. A. Dalrymple, and M. A. Losada (1998). Wave-induced mean flows in vertical rubble mound structures. Coastal Eng. 35: 251-281. Lynett, P. J., P. L.-F. Liu, I. J. Losada, and C. Vidal (2000). Solitary wave interaction with porous brealwaters. J. Wtrway., Port, Coast, and Oc. Eng. ASCE 126(6): 314-322. Madsen, O. S. (1974). Wave transmission through porous structures. Journal of Waterways, Harbours, Coastal Engineering Division. ASCE 100(WW3): 168-188. Mase, H. and K. Takeba (1994). Bragg scattering of waves over porous rippled bed. In: Proc. 84th ICCE, Kobe. 635-649. Massel, S. R. and C. C. Mei (1977). Transmission of random wind waves through perforated or porous breakwaters. Coastal Eng. 1(1): 63-78. Massel, S. R. and P. Butowski (1980). Wind waves transmission through porous breakwatwers. Proc. 17th Coastal Engineering Conference, ASCE. New York. 333-346. Mclver, P. (1998). The dispersion relation and eigenfunction expansions for water waves in a porous structure. Journal of Engineering Mathematics 34: 319-334.
200
/. J. Losada
Mendez, F., I. J. Losada, and M. A. Losada (2001). Wave-induced mean magnitudes in permeable submerged breakwaters. J. Wtrway., Port, Coast, and Oc. Eng. ASCE 127(1): 1-9. Polubarinova-Kochina, P. Y. (1962). Theory of Groundwater Movement. Princeton University Press, Princeton, N. J. Rivero, F. J., A. S.-Arcilla, X. Gironella, and A. Corrons (1998). Large-scale hydrodynamic experiments in submerged breakwaters. Proc. Coastal Dynamics'97, ASCE. Virginia. 754-762. Rojanakamthorn, S., M. Isobe, and A. Watanabe (1989). A mathematical model of wave transformation over a submerged breakwater. Coastal Engineering in Japan, JSCE, 32(2): 209-234. Rojanakamthorn, S., M. Isobe, and A. Watanabe (1990). Modeling of wave transformation on submerged breakwater. Proc. 22nd Coastal Engineering Conference, ASCE. New York. 1060-1073. Rodi, W. (1980). Turbulence models and their application in hydraulics — a stateof-the-art review. IAHR Publication. Shih, T.-H., J. Zhu, and J. L. Lumley (1996). Calculation of wall-bounded complex flows and free shear flows. Int. J. Numer. Meth. Fluids 23: 1133-1144. Smith, G. (1991). Comparison of Stationary and Oscillatory Flow Through Porous Media. M.Sc. Thesis. Queens University. Canada. Sollitt, C. K. and R. H. Cross (1972). Wave transmission through permeable breakwaters. Proc. 13th. Coastal Engineering Conference, ASCE. New York. 1827-1846. Sulisz, W. (1985). Wave reflection and transmission at permeable breakwaters of arbitrary cross-section. Coastal Eng. 9: 371-386. Troch, P. and J. de Rouck (1998). Development of a two-dimensional numerical wave flume for wave interaction with rubble mound breakwaters. Proc. 26th. International Conference on Coastal Engineering, ASCE. Reston, Va. 1638-1846. Vidal, C , M. A. Losada, R. Medina, J. Rubio (1988). Solitary wave transmission through porous breakwater. Proc. 21st International Coastal Engineering Conference 1073-1083. Ward, J. C. (1964). Turbulent flow in porous medium. Journal of the Hydraulics Division, ASCE 90, HY5. 1-12. Wu, T. Y. (1981). Long waves in ocean and coastal waters. J. Eng. Mech. Div. 107: 501-522. Yu, X. (1995). Diffraction of water waves by porous breakwaters. J. Wtrway., Port, Coast, and Oc. Eng. ASCE 121 (6): 275-282. Yu, X. and H. Togoshi (1996). Combined diffraction and transmission of water waves around a porous breakwater gap. Proc. 25th Int. Conf. Coastal Eng., ASCE 2063-2076.
Recent Advances in the Modeling of Wave
201
List of Symbols / p g po ho ud K ap Kp (m 2 ) Re = ucDc/v uc Dc v bp ne ap (3P cp 7 fl fff // KC u* p* Ui u\ u\ Cj s / u> h b $ fo
= hydraulic gradient = fluid density — gravitational acceleration = the effective pressure = the vertical distance from the selected datum = discharge velocity = l/ap (m/s) permeability coefficient = empirical coefficient = intrinsic permeability = Critical Reynolds number = characteristic discharge velocity = characteristic length scale of the porous media = molecular viscosity = empirical coefficient with dimension (s 2 /m 2 ) = the porosity = nondimensional coefficient = nondimensional coefficient — empirical coefficient (s 2 /m) = nondimensional coefficient accounting for the added mass = resistance force due to laminar flow = resistance force due to turbulent flow = resistance force due to inertia = Keulegan-Carpenter number = ith component of the instantaneous velocity in the pores = instantaneous effective pressure = ith component of the seepage velocity = ith component of the spatial perturbation = ith component of time perturbation = dimensionless turbulent coefficient = co-factor accounting for added mass = linearized friction coefficient = wave frequency = water depth = porous structure width = velocity potential = energy dissipation function associated with wave breaking
202
I. J. Losada
To = T.R + iTi = complex wavenumber in the porous medium C = free surface displacement Hrms = mean-root-square wave height Tp = peak period hef = effective water depth 4> = depth-averaged potential u = depth-averaged potential tp = depth-averaged piezometric head 5ij = Kronecker delta k = turbulent kinetic energy £ = dissipation rate of turbulent kinetic energy Ui = seepage current Ui = seepage oscillatory flow U* = spatial fluctuation of the current U\ = temporal fluctuation of the current u\ = spatial fluctuation of the oscillatory flow u\ = temporal fluctuation of the oscillatory flow Sij = components of the radiation stress due seepage magnitudes Sfj = components of the radiation stress due to the spatial fluctuations Sjj = components of the radiation stress due to the temporal fluctuations
D E S C R I P T I V E H Y D R O D Y N A M I C S OF LOCK-EXCHANGE FLOWS
HARRY YEH and KIYOSHI WADA Salt-water gravity currents and internal bores are created in a horizontal flume with a lock-exchange device, i.e., by lifting a partition gate that initially separates fresh water from salt water. Using laser-induced fluorescein-dye illumination in the laboratory, qualitative characteristics and behaviors of the flows are examined. In the second set of the experiments, gravity currents and internal bores are first generated in a channel of finite breadth to establish its quasi-steady condition. Then, the established current is let off to spread from the end of the channel to open environment. The spreading flow pattern and mixing around the end of the channel is examined and interpreted based on vortex dynamics.
1. Introduction Gravity currents are gravity-driven flows by the fluid-density difference from that of its surroundings. When a similar flow advances into a quiescent twolayer fluid of which the thinner layer has the same fluid density as the advancing current and distinct wave breaking is formed at the leading wave, the flow is called an internal bore. Gravity currents and internal bores are investigated in a horizontal laboratory tank with a lock-exchange device, i.e., by lifting a partition gate that initially separates fresh water from salt water. Many researchers have investigated the fundamental flow characteristics of gravity currents and internal bores. What we wish to present here is the clarification of features and mixing mechanisms associated with gravity-current and internal-bore phenomena, and the differences between the two flows. Furthermore, observing the sudden expansion of the current to open environment, the three-dimensional flow patterns and resulting mixing mechanisms are described. Features and hydrodynamics of such flow phenomena are discussed particularly in terms of vortex dynamics. Based on the assumptions of an inviscid fluid and a two-dimensional flow field between a pair of infinite parallel horizontal plates (no free surface 203
204
H. Yeh & K. Wada
involved), Benjamin (1968) established the lowest-order theory for gravity currents flowing on a horizontal bed. Combining the depth-integrated linearmomentum equation and continuity, Benjamin derived that:
^=(2^i)(l-§) g'h
(1)
(1 + 1)
where, as shown in Fig. 1, h is the current depth, d is the total flow depth, which is the spacing between the two parallel horizontal plates, U is the speed of the front, F is the densimetric Froude number, g' = (p2 — Pi)d/P\ is the buoyant acceleration, where p\ and p2 are the fluid densities of the ambient and the current respectively, and g is the acceleration of gravity. Benjamin (1968) ironically found that the inviscid theory breaks down except for in a very special case; the only possible solution to satisfy Eq. (1) and the Bernoulli theorem is h/d — 1/2 and F = 2 - 1 / 2 . The steady-state condition of h/d > 1/2 is not possible, while the condition of h/d < 1/2 requires energy dissipation by wave breaking. Benjamin's solution of this very special case implies that energy dissipation associated with turbulence must play an essential role in determining the dynamics of gravity currents. Not only for that, turbulence that intrinsically generated must cause mixing of two fluids that make up the gravity currents. Figure 1 summarizes the features of an inviscid gravity current. Several extensions of Benjamin's work by incorporating energy dissipation are proposed by Wilkinson and Wood (1972), Chu and Baddour (1977), Wood and Simpson (1984), Denton (1990), and Klemp et al. (1994). These models '
'
*
•>
'
'
£
S *
£
£.
-»
-*
'
'
'
1
h /7I/3 •
/
/
i /—7—7—7—7—7—7—7-7-7—7—?—r
Fig. 1. A schematic of inviscid gravity current that is equivalent to the model made by Benjamin (1968). For the inviscid model, the height of gravity current is one-half of the channel height and the leading edge contacts at 7I"/3 with the bed, which is basically the Stokes 120° corner flow.
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
205
are based on the mixing-layer formation. The model by Wood and Simpson assumes that energy dissipation occurs only in the lower layer, whereas Denton's model assumes energy dissipation in the upper layer. Using a perturbation analysis, Jirka and Arita (1987) showed that the head formation of a gravity current becomes a density wedge (without a head formation) in the case of momentum deficit relative to the inviscid case. A steady gravity current can be maintained only if there is a momentum surplus. In his book, Simpson (1987) summarized that, as a gravity current advances, a characteristic head is formed at the leading edge of the gravity current, which is approximately twice as deep as the following flow, H « 2/i, as depicted in Fig. 2. The characteristics of the head are considered to control the entire flow behavior, e.g., mixing, velocity of the advancement, and its profile. The head profile is sensitive to the Reynolds number (R = Uh/is), the densimetric Froude number (F = U/y/g'h), and the ambient flow conditions (Note that in the Reynolds number, v is the kinematic viscosity of the fluid). There are according to Simpson (1987) two types of instabilities that are responsible for mixing associated with gravity currents: (1) billows which roll up in the region of velocity shear above the advancing front and (2) a complex shifting pattern of "lobes and clefts" located on the face of the head. Simpson suggested that the billow formation on the upper surface of the head is similar to the Kelvin-Helmholtz instability. He also indicated that the formation of "lobes and clefts" develops from the lighter fluid that is overrun by the gravity current's foremost leading edge. The foremost leading edge of the current is located slightly above the bed; approximately 1/8 of the total height of the head. Hence, some ambient fresh water is trapped under the head and
Fig. 2. A sketch of a typical gravity current in a laboratory channel. The forefront nose is approximately 1/8 of the total height of the head H. Because of this feature, the lighter fluid can be entrapped under the nose, which causes the formation of flow pattern "lobes and clefts" on the front face of the head. The broken line indicates the location of laser sheet used for Fig. 6.
206
H. Yeh & K. Wada
entrained into the current from the bottom. This trapped lighter fluid is advected upward through the denser fluid due to buoyancy to form a complicated shifting pattern of "lobes and clefts" on the front face of the head. The majority of previous work was to investigate the phenomena in a two-dimensional channel, i.e., the mean variations in the horizontal direction perpendicular to the current direction were neglected. Investigation remains insufficient for the mixing processes of two-layer flows disturbed by "vertically intruding" objects which evidently cause three-dimensional disturbances. Such a mixing process occurs, for example, at irregular estuarine side boundaries, jetties, piles and the like, and appears to be effective and efficient in mixing. It is noted that locally created vertical mixing often causes horizontal density gradients which may drive a transverse circulation in the entire water body (Fischer, et al, 1979; Maxworthy and Monismith, 1988); this process could potentially be dominant for mixing in some estuaries. In this paper, we first discuss fundamental features of gravity currents and internal bores. Then, three-dimensional mixing characteristics associated with sudden expansion of gravity current and internal bore in a two-dimensional channel into open environment are examined. Strong three-dimensional disturbances are observed with flow visualization and the dominant flow characteristics are identified. Flow behaviors associated with the expansion should be similar to the disturbance caused by "vertically intruding" objects. 2. Experimental Facilities Two different experimental apparatuses were used to examine a variety of flow behaviors associated with gravity currents and internal bores. First, traditional experiments for lock exchange flows were performed in a 16.2 in long, 0.61 m wide, and 0.45 m deep tank. The tank was initially divided into two separate chambers with the aluminum gate: The front chamber is 8.9 m long and the back chamber is 7.3 m long. For the gravity current experiments, uniformly mixed salt water filled the back chamber, while fresh water filled the front chamber. For the internal bore experiments, a thin layer of salt water was placed beneath the fresh water in the front chamber, and the fluid density of the thin layer was identical to that of the salt water in the back chamber. A sharp interface was established by introducing salt water slowly through the diffuser along the bottom against the side wall. Throughout a series of experiments, the total water depths of both front and back chambers are set at 36 cm. A gravity current or an internal bore
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
207
was created as the lock-exchange flow by lifting the aforementioned aluminum gate vertically to 20 cm from the bed. This partial gate opening is a similar generation scheme as that used by Wood and Simpson (1984), which minimizes free-surface disturbance created at the gate and controls the current depth to approximately 9 cm. Horizontal and vertical laser sheets were employed to resolve the flow behavior of gravity currents and internal bores. A 4-watt Argon-ion laser beam is shone onto the oscillating mirror, which produces a laser sheet. The laser sheet illuminates the fluid dyed with disodium fluorescein (C20H10O5N2) in the flow. The dye is previously mixed with the saltwater. The Argon-ion laser operates in the 457.9-528.7 nm wavelength range, i.e., the green spectrum. Fluorescein dye is excited by the green spectrum and becomes a brilliant yellowgreen. This flow visualization technique is often referred to as the laser-induced fluorescence (LIF) technique. Observations by the LIF technique were made in the front-chamber area 4.5 m downstream of the gate, approximately 50 times the average saltwater flow depth. All the transient disturbances caused by the gate motion should have sufficiently subsided in the area of the observation and measurements. It is also noted that the length of the back chamber is long enough so that flows in the observation area are not disturbed by the reflection from the back-chamber end wall. The conductivity-temperature instrument was used which measures the electrical conductivity and temperature of a solution. An estimate of the spatial resolution is a 1 mm diameter sphere. There is 1 mm separation between the conductivity sensor and the thermistor bead, which implies that the two sensors are reading nearly the same parcel of fluid. The traversing system was used at its maximum traveling velocity in order to obtain an "instantaneous" profile of the flow. At the maximum velocity, the traversing system takes about 2 s to travel down and up through the water column. With the data-scanning rate of 400 scans/s, the spatial resolution of the probe is approximately 1 mm. More detailed discussions of this laboratory apparatus are given in Grandinetti (1992). The second set of experiments was designed to examine the flow behaviors and characteristics associated with flow expansion of established gravity currents to a much wider reservoir. This series of experiments was performed in a 3.0 m long, 1.2 m wide, and 0.9 m high glass-panel water tank. Within the tank, a false partition wall was placed parallel to the tank sidewall to form a
208
H. Yeh & K. Wada
(a) Elevation View (A-A Section) .
.
.
.
.
;
•
•
'
•
-
'
-
'
.
Laser\
vyZtCvyA Q I » A I > fa,
Scanner
W.
(Unit:cm)
Gate
I Mirror |
J8
V
Iho
f^Laser \ -/~jsheet^\.
t
z ' Fresh Water
J Salt Water [
II •o
,
t
X
300
(b) Plane View CM
180
i
Partition Wall
II
n
9°
»
Not to scale Fig. 3. Schematic views of the experimental set up: (a) elevation view and (b) plan view. Note that the origin of a Cartesian coordinate system with the right-hand rule is taken at the end of the intersection between the partition wall and the channel bed.
0.22 m wide and 1.80 m long confinement as shown in Fig. 3. Gravity currents and internal bores were also generated by the lock-exchange mechanism, i.e., by lifting the gate to form a flow pattern guided by the confinement. The generated currents were then expanded into open environment at the end of the confinement. Note that the expansion took place on only one side of the channel. In order to describe the locations and directions, we define the coordinate system as shown in Fig. 3. The origin is taken at the bottom of the channel at the tip of the partition wall. Taking Cartesian coordinates from the origin with the right-hand rule, the x-axis points horizontally in the direction extending the partially confined partition wall, the y-axis points horizontally in the direction toward the opposite wall and the z-axis points upward. We conveniently term the partition wall (at y = 0) the "right" wall and the fully extended wall (at y = 0.22 m) the "left" wall. The tip of the partition is identified by the location (x = 0,y = 0). The directions of fluid rotation are described based
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
209
on the right-hand rule of the coordinate system. We also set the time origin (t = 0 s) when the leading edge of the current passes at the position x = 0. The resulting flow patterns were visualized using the laser-induced fluorescent (LIF) technique. 3. Basis for Interpretations of Flow Images The primary interest of hydrodynamics associated with gravity currents and internal bores is their ability to generate turbulence at the interface and subsequently to induce mixing. Hence, it is necessary to review briefly the mechanisms of turbulence generation. Turbulence is rotational flows although the inverse may not be always true: There are many examples for laminar rotational flows. On the other hand, there is no turbulent flow that is irrotational. On this basis, there are two possibilities to generate turbulence associated with gravity currents. First, the fluid is initially rotational (vortical) by some reason, and turbulence is generated by stretching and bending of pre-existing vortex tubes. Under the condition of lock exchange flows, both fluids (saline and fresh water) separated by the lock are considered to be quiescent and initially irrotational. The other possibility is that fluid rotation is created within the fluid domain during its propagation. The fluid rotation then leads to turbulence by stretching and bending the created vortex tubes. For Newtonian mechanics, there must be a force to accelerate a fluid parcel, which is the statement of the conservation of linear momentum. Likewise, the conservation of angular momentum requires torque to create rotation of a fluid parcel when it is initially irrotational. In other words, if there is no torque, then fluid remains irrotational. From this fundamental point of view, we pay careful attention to torque that arises from lock-exchange flows. Fluid rotation can be conveniently measured by flow circulation T, which is defined by:
T= fudx= Jc
fuj-dA,
(2)
Js
where u is the fluid velocity, u> is the vorticity (curl u), x is the position vector of a closed integration contour c, and A is the area vector of the surface s whose boundary is a single closed curve c. The rate of change in circulation T following a fluid parcel that make up the curve c can be found to be (e.g., see Lighthill, 1987):
ft -Liix)-*1-!.&•*')•*•
(3)
210
H. Yeh & K. Wada
where r^- is the stress tensor, p is the fluid density, and D/Dt is the material derivative. In Eq. (3), we used the Cauchy equation with the conservative body force (e.g., see Serrin, 1959). For incompressible Newtonian fluids with uniform viscosity, using the Stokes theorem, Eq. (3) can be written as: DT Dt
j
— (Vp x Vp) + - V2w - — (Vp x pV 2 u)
s
|_p' '
'
•'
p
(f
dA,
(4)
where p is the pressure and fi is the dynamic viscosity of the fluid. The second term on the right-hand side of Eq. (4) represents the diffusion of vorticity from fluid parcels adjacent to the boundary of the integration surface s. Hence, this does not represent vorticity creation within the fluid domain, but represents vorticity transfer from the surroundings (or the boundaries) due to viscous diffusion. In the case of inviscid fluids, flow circulation can be produced within the fluid domain whenever the fluid is displaced from a state in which the pressure gradient Vp and the density gradient Vp are parallel, i.e., Eq. (4) can be modified to be: f
= /
s
> P x V p
W
A .
(5)
This is often called Bjerknes' theorem (see, for example, Lamb, 1932, Article 166-a). Equation (5) represents the time rate of change in circulation by torque acting on a fluid parcel: This torque is often termed "baroclinic torque": Physical interpretation of baroclinic torque was discussed in, for example, Yeh (1995). The last term in Eq. (4) represents net viscous force acting upon a fluid particle. It is emphasized that, under the influence of a conservative body-force field, rotational motion can be produced from an initially irrotational state only by two mechanisms: Baroclinic torque and the interaction of density gradient and viscous shear force, and both mechanisms require the existence of density gradients. The latter rotation creation mechanism is termed as "viscous-shear torque" (Yeh, 1991, 1995). If we further assume the fluid to be inviscid and homogeneous, then Eq. (5) is reduced to the Kelvin theorem:
Equation (6) states that flow circulation T around a closed curve c moving with the fluid remains constant.
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
211
Now at this point, it is evident that fluid rotation can be created within the fluid domain only by the action of baroclinic torque or/and viscous shear torque. Consider a typical flow pattern resulted from a lock-exchange flow moving with the velocity U. To compare the two mechanisms of fluid-rotation creation associated with lock exchange flows, first, we note that the gradient of fluid density Vp always points in the direction normal to the interface. Hence, the order of magnitude of the ratio of viscous shear torque l/p2Vp x /xV 2 u to baroclinic torque l/p2Vp x Vp along the frontal interface is: /uV
2
Ma_b Va_6p '
l
>
where the subscript a-b indicates the component in the direction along the interface a-b as shown in Fig. 4.
*
A /—7—7
7
7—7—7—7
7—7
7
T—f
T
L Fig. 4. A definition sketch for the front of an internal bore. Fluid rotation is induced at the front due to the presence of sharp density gradient at the interface by baroclinic torque.
Since the divergence of viscous-shear stresses, /xV2u represents a net viscous force per unit volume acting on a fluid parcel and the force must be continuous across the salt-and-fresh water interface, the value of /iV 2 u at the interface can be evaluated either in the fresh or salt water domain. We estimate this value based on the assumption of thin-boundary-layer analogy in the domain of fresh water; the velocity outside of the boundary layer can be considered to be small at the front in the stationary reference frame (the same is not true in the salt-water domain and the thin-boundary layer analogy cannot be justified. Consequently, it is difficult to estimate the values of /LtV2u by using the saltwater velocity field). The Laplacian of the water velocity along the interface
212
H. Yeh & K. Wada
can be estimated by: 72 .
_. ,.d2Ua-b
_,^f
pU
/ i V ^ a _ 6 « P~-^ » O (-^— ) , (8) a; \icosa/ where u a _b « U/ cos a is the water velocity parallel to the interface in the coordinates moving with the current, a is the angle of the front face from the horizontal, £ points in the direction normal to the interface as shown in Fig. 4, U is the propagation speed, and we used the boundary-layer thickness to be 0{y/vi) where v is the kinematic viscosity of the water {y = p/p). Along the frontal interface (a-b in Fig. 4), the time scale t can be estimated by t « L/U in which L is the length of the front as shown in Fig. 4, hence Eq. (8) can be written as: pU2 / J W - 6 » O -f . (9) \ L cos a J The gradient of pressure field along the frontal interface may be estimated based on the assumption of hydrostatic pressure field, which is: dri Va-bpK, pg—tt pg sin a, (10) where r] is the elevation of the interface and £ points in the direction along the interface a-b as shown in Fig. 4. The ratio of baroclinic torque to viscous shear torque is then written as: MW-6 Va-bp
W
U2 U2 W gL sin a cos a gh* cos2 a '
(11)
where h* is the height of the front, i.e., h* m L tan a (see Fig. 4). It is emphasized that g is the acceleration of gravity but not the reduced gravity g ^ g' = gAp/p, that often arises from the analyses of stratified flows. Our estimate in Eq. (11) demonstrates that with the exception of forced flows such as a jet by maintaining the interface slope very small, the ratio of viscous shear torque to baroclinic torque is extremely small. For example, the front face of typical lock-exchange flow with Ap/p = 0.002, p = 1,000 kg/m 3 and fj, = 0.001 kg/m-s, the densimetric Froude number U/yfgH is approximately unity, hence U2/gh w U2/gh* is 0.002. If the interface slope is less than 25 degrees, the baroclinic torque is more than 400 times greater than the viscous shear torque. Even without the aforementioned quantitative estimate, it can be argued that the magnitude of viscous shear torque is usually small at the interface. It
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
213
is because the inflection point of the velocity profile of a two-layer flow is often located at or very close to the density interface. At the inflection point where the curvature of the velocity profile vanishes, i.e., V 2 u = 0, hence, there is no viscous shear torque. It is now fair to say that the dominant mechanism for the creation of fluid rotation for free lock-exchange flow is baroclinic torque at the salt-fresh water interface. The flow characteristics and behaviors that we will present in this article will be primarily interpreted based on this concept. 4. F e a t u r e s of G r a v i t y C u r r e n t s Figures 5 and 6 show our laser-induced-fluorescence flow images in a longitudinal vertical plane at the center of the tank and in a horizontal plane 10.5 cm above the bed respectively. For the images shown in Fig. 6, the location of the laser illumination plane is depicted in Fig. 2. Only the salt water (p2 — 1002 kg/m 3 ) was dyed with fluorescein and appears as bright regions having been illuminated by the Argon-ion laser sheet. The Reynolds number and the densimetric Froude number of this flow are R = 1450 and F = 0.995 respectively. The flow images in Fig. 5 show the formation of billows on the front face; these small-scale billows are found to be three-dimensional. This can be verified in Fig. 6 where the front's pattern in the horizontal plane is irregular. This is different from the features of a typical Kelvin-Helmholtz instability in a plain shear flow that is, for example, created by a splitter plate; the billows
(a)
(b)
Fig. 5. Gravity-current flow structure in the longitudinal vertical plane, the Reynolds number R = 1450 and the internal Froude number F = 0.995. (a) Laser-induced fluorescence shows the irregular front face and the ascension of entrapped fresh water and (b) the inverted roll-up under the nose of the leading front. Time interval between (a) and (b) is 5.6 s. Note that the faint horizontal band inside the gravity current shown in the photographs is the reflection from the diffuser.
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
215
associated with the Kelvin-Helmholtz instability are at least initially twodimensional, i.e., the roll-up features of billows are long-crested. Instead, the three-dimensional front formation observed here is consistent with the "lobes and clefts" pattern indicated by Simpson (1987). There are several intriguing features in Figs. 5 and 6 that need to be addressed. The underside entrainment of the lighter fluid at the front of the gravity current is clearly seen in Fig. 5. In Fig. 5(a), the dark streaks being flared near the very front along the bed show the ascension of entrapped fresh water due to its buoyancy. Figure 5(b) shows an underside entrainment; the inverted roll-up and entrainment of the ambient fluid under the "nose" can be identified. The leading edge is irregular as seen in Fig. 6, including the formation of counter-rotating, mushroom-shaped eddies (Fig. 6(a)), which are presumably associated with the formation of U-shaped vortex loops on the front face. In the flow following the front (see Fig. 6(b)), the pattern is irregular, although it appears to be an alternating pattern of salt and fresh water rows in the propagation direction indicating periodic formation of large-size eddies. 5. Features of Internal Bores Our experimental results in Fig. 7 show that the flow characteristics of the gravity current and internal bore are significantly different from each other, even when the pre-existing front layer of the denser fluid for the internal bore is thin (0.5 cm for Fig. 7(b), i.e., approximately 1.4 percent of the total depth while Figs. 7(c) and 7(d) are 1 cm and 2 cm respectively). Note that the Reynolds number and the densimetric Froude number of the flows shown in Fig. 7 are (a) R = 1450 and F = 0.995 and (b) R = 1440 and F = 1.00, (c) R = 1730 and F = 1.29 and (d) R = 1230 and F = 0.861 respectively. Unlike a gravity current, the front face of an internal bore is smooth, and turbulence is generated at the rear side of the head. Figure 7 shows that the thinner the initial front layer, the steeper the front face and the earlier the formation of billow roll-ups. Contrary to the case of gravity currents, the billow formation of an internal bore resembles that of the Kelvin-Helmholtz instability, which appears to be two-dimensional at least initially, i.e., the flow pattern is uniform in the direction transverse to the flow. The flow pattern of the internal bore (the case shown in Fig. 7(d)) is further examined by introducing a small red dye (Rhodamine) patch in the initial salt-water layer as shown in Fig. 8. It is evident from the time sequence of
218
H. Yeh & K. Wada
photographs that the initially undisturbed fluid of the thin layer moves along the front face (saline and fresh water interface) and the bottom boundary; i.e., it looks like the dyed fluid is coating the advancing bore head. The distinct differences in the features between gravity currents and internal bores might be explained by Simpson's (1987) "lobes and clefts" formation at a gravity current's foremost leading edge. Such formations are not possible in an internal bore because ambient fresh water cannot be trapped along the bottom where the denser fluid already occupies. Further explanation can be given based on the vorticity creation mechanisms given by Eq. (4). Recall that there are only two mechanisms to create fluid rotation within the fluid domain (excluding at solid boundary surfaces). These are baroclinic torque (the first integrand term in Eq. (4)) and viscous-shear torque (the last integrand term) both of which require the presence of density gradient. It is emphasized that the second integrand in Eq. (4) represents the transfer of fluid rotation by vorticity diffusion from fluid parcels adjacent to the boundary of the integration surface, and does not represent creation of fluid rotation within the fluid domain. Also note that it was shown in Eq. (11) by an order-of-magnitude analysis that viscous-shear torque plays an insignificant role compared with the role of baroclinic torque. By following the red-dye fluid patch shown in Fig. 8, the leading bore front consists of fluid parcels previously located in the quiescent thin layer in front of the bore. Hence, in the case of the internal bore, the fluid parcels along the interface are initially quiescent and irrotational. Fluid rotation must be created at the interface by baroclinic torque: The pressure gradient must be close to that of hydrostatic condition while the density gradient is normal to the interface as shown in Fig. 9. Once fluid rotation was created by baroclinic torque, the rotationality is advected with the fluid motion (Helmholtz's theorem) and diffuses by viscous effects. This property manifests itself in the formation of billows behind the ridge of the bore front. On the other hand, for a gravity current, the fluid parcels along the interface are advected from inside of the advancing current as shown in Fig. 9(a). The fluid within the current is already vortical due to turbulence induced by (a) wave breaking behind the head, (b) entrapped fresh water under the leading edge or nose (see Fig. 5), and/or (c) the bottom (no-slip) boundary condition. Hence, fluid parcels along the interface are originally vortical. This vortically perturbed flow is affected further by the creation of fluid rotation by baroclinic torque at the front interface, which results in three-dimensional
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
219
Vp Vp
-?—s /
-7—7—v—7—7—r-
s s >
(a) Vp
Vp
/ /
-r—r
(b) Fig. 9. A sketch of flow patterns at the front of (a) gravity current and (b) internal bore. Fluid-rotation is induced along the front face by baroclinic torque due to misalignment of the density gradient Vp that is perpendicular to the interface and the pressure gradient Vp that is approximately vertical for hydrostatic condition.
and complicated patterns at the interface of the head as shown in Figs. 5 and 6. The nature of inherently vortical flow at the gravity-current head can be an explanation for "lobe-and-cleft" formations as well as the reason for three-dimensional small-scale billow formations. It appears in Figs. 7 and 8 that the vortex formation behind the head of the internal bore resembles that of a separation eddy associated with sudden flow expansion of the fresh water along a fictitious saline wall (i.e., the saline-fresh water interface). Furthermore, roll-ups form a periodic pattern of turbulent regions at somewhat constant intervals. The periodic features of turbulence patches might be related to the intermittent nature of the flow. It is conjectured that turbulence patches are formed by the "generation-advection" cycle of the roll-up formation. As soon as the roll-up is formed, the large vortex of the roll-up is advected behind the head, then the next roll-up is created, and this process is repeated. Yeh and Mok (1990) demonstrated that turbulence-patch formations behind bores at the air-water interface also result from the generation-advection cycle. In order to test this hypothesis, the periodic turbulence-patch formation is quantified by the frequency of its
220
H. Yeh & K. Wada
generation; the generation frequency can be measured directly from the video images of our experiments. This measured frequency should be related to the frequency associated with the excursion time of a fluid parcel traveling around the vortex. Assuming the vortex size 5 (say the diameter of the orbital motion) is comparable with the difference in height between the head and the current depth behind the head, and using the vortex-velocity scale to be the bore propagation velocity U, then the time scale for roll-up is found to be Tr w irS/U. The vortex size S as indicated in the sketch shown in Fig. 9(b) (the actual length scale of the vortex is somewhat smaller than this estimation as seen in Figs. 7 and 8) and the propagation speed U were measured directly from the time-lapse flow images. Computed frequencies l/Tr shown in Fig. 10 are in a very good agreement with the measured frequency from the video images; note that the error bars in the figure represent the 90% confidence limit for the frequency measurements. The excellent agreement in Fig. 10 supports the conjecture that the periodic behavior of the turbulence patches
1.0
•o 0.8 o> t3 'S a>
a. 0.6 f? & 0.4 c ID 3 CT 0)
*
0.2
0.0 1
0.0
I
T
I
I
0.2 0.4 0.6 0.8 frequency (Hz) measured
I
1.0
Fig. 10. Predicted and measured generation frequencies of the large-scale eddies: O, gravity current (Fig. 7(a)); x, internal bore with ho = 0.5 cm (Fig. 7(b)); A, internal bore with ho = 1.0 cm (Fig. 7(c)); o, internal bore with ho = 2.0 cm (Fig. 7(d)). The error bars represent 90% confidence limits. The prediction is based on the hypothesis that each eddy is formed and advected at the rate of the excursion time for a fluid parcel travelling around the roll-up.
224
H. Yeh & K. Wada
the three-dimensional variations appears to be plausible for the present case. To see this more clearly, consider the time scales of the vortex motion and the buoyancy effect. The buoyancy time scale TJ, can be estimated as the traversing time for a fluid-parcel to ascend through the vortex length scale S by the buoyancy force: Tb « y/5/g'. Assuming the vortex size 5 is comparable with the difference in height between the head and the current depth behind the head (see Fig. 9(b)) and using the buoyant acceleration g' = (p2 — Pi)g/pi, the time scale for the buoyancy effect is found to be T(, w 1.2 s for the case shown in Fig. 12. The roll-up time scale can be estimated as before: An excursion time for a fluid parcel to travel around the vortex. Using the vortex velocity scale to be the bore propagation velocity U, the time scale for roll-up is found to be Tr « nS/U = 3.8 s. This rough estimate demonstrates that the order of computed magnitudes of the buoyancy-effect time scale and of the vortex-motion time scale is comparable, which supports the explanation for three-dimensional variations discussed the above. Rather than the manifestation of longitudinal (streamwise) vortex formations in the Kelvin-Helmholtz instability, the gravitational instability is more likely the one that causes the transverse perturbation of vortices that appears in Fig. 12. Typical density-profile data of an internal bore are presented in Fig. 13. The results in Fig. 13(a) show the mixing phase caused by overturning billows; the small-scale density inversions in the figures can be interpreted as the presence of small active eddies. Approximately 10 s after the passage of the head, these small-scale density inversions are diminished (Fig. 13(b)). At this point, the region of overturning billows is considered to be mixed, i.e., appreciable fluid mixing has already taken place in the vicinity of the head. The results appear to be consistent with our discussion made earlier, indicating that fluid mixing at the head is immediate and efficient in comparison to mixing in the current far behind the head. In fact, Figs. 13(c) and 13(d) show no density inversion that means no significant fluid entrainment being taken place. Note that two separate mixing processes are represented by the appearance of two distinguishable gradients in the density profile plot (Fig. 13(d)). The upper portion of the profile has a uniform density gradient caused by largescale mixing of overturning billows. As mentioned, this mixing process takes place directly behind the head. In the second gradient, the mixing process is attributed to the continual shear flow created by the fast-moving dense water under the nearly stationary mixed-fluid region. This fast-moving fluid layer is vortical due to the boundary-layer effect, which gradually erodes the layer, as
Descriptive
Hydrodynamics
of Lock-Exchange
1000
1002
Flows
225
is-
le-
v«' V* s'
n997
998
999
1001
1003
3
p (kg/m ) (a)
15
3. .c -a
10
S
*.
5-
^ 0 997
998
999
1000
1001
1002 1003
p (kg/m3) (b) Fig. 13. Time sequence of the density profile for an internal bore shown in Fig. 8, R = 1230, F = 0.861; (a) t = t 0 , (b) t = t 0 + 11 s., (c) t = t 0 + 21 s. and (d) t = to + 32 s. The small-scale density inversions shown in (a) due to active eddies are diminished in 11 s. as shown in (b). This indicates that appreciable fluid mixing took place quickly. Thereafter, the density profile remains relatively unchanged.
226
H. Yeh & K. Wada
15
V ?
io-
\
.
*-..
a. T3
•
^
0
997
998
999
1000
1001
1002 1003
3
p (kg/m ) (c)
15
v..
-4—'
a* o
"N
5^ — 1
997
1
998
1
1
999
1
1
1000
1
1
1001
p (kg/m3) (d) Fig. 13.
{Continued)
1
I
1
1
1002 1003
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
227
do the combined effects of shear instability and small-scale eddies within the lower layer fluid. It is also emphasized that the immediate mixing appears to take place for the fluid initially placed in front of the internal bore, while the fluid in the core of the internal bore tends to remain unmixed as shown in Fig. 11. 6. Flow E x p a n s i o n of Gravity Currents and Internal Bores Three-dimensional mixing characteristics associated with sudden expansion of a gravity current and an internal bore into open environment are examined using the second experimental apparatus (see Fig. 3). In order to demonstrate the effects of expansion, we first repeat the experiments in the second apparatus for the gravity current and internal bore in the confined narrow channel by blocking the expansion. It is evident in Fig. 14 that the similar flow patterns are obtained as those shown in Fig. 7. Note that the Reynolds number and the densimetric Froude number of the flows shown in Fig. 14 are (a) R = 1700 and F = 0.73 for the gravity current and (b) R = 1800 and F = 0.78 for the internal bore respectively. Effects of spreading of gravity current and internal bore can been seen clearly by comparing the longitudinal flow profiles in Figs. 14 and 15. Figure 15 shows the flow patterns at y/d = 0.17 where d is the total flow depth (0.15 m) and the location of x — 0 is indicated with the mark T in the figure (see Fig. 3 for the coordinates used in the discussion). The leading face of the gravity current has a complex flow pattern, while the internal bore has a smooth front face followed by the formation of roll up. While the difference in the head characteristics between the gravity current and the internal
(a)
•trKv
(b) Fig. 14. Longitudinal profiles of (a) gravity current (R = 1700, F = 0.73) and (b) internal bore (R = 1800, F = 0.78) in a channel with a uniform width. Note that the features are similar to those shown in Fig. 7.
228
H. Yeh & K. Wada f
* * * /JTP?
*
T (a)
(b) Fig. 15. Longitudinal profiles of (a) gravity current (R = 1700, F = 0.73) and (b) internal bore (R = 1800, F = 0.78) along the vertical plane at y/d = 0.17 when they spread out into open environment. Note that the location of the end of the false partition wall is marked by V in (b).
bore remains the same, the leading head heights are substantially reduced due to spreading. Behind the leading head, the flow depth is grossly depressed and then the mixed fluid is lifted up to the free surface at the location near x = 0 (Fig. 15). The variations of the depressed flow depth behind the head are plotted in Fig. 16 for the gravity current (the results for internal bores are similar and not shown here). The depth of the two-dimensional gravity current in the confined channel remains constant at hm\n/d « 0.4. On the other hand, once the gravity current spreads into open environment (x > 0), the depth behind the head decreases substantially to hmjn/d w 0.1. The reduction of the depth occurs first near the confined wall (y m 0) and the effect propagates toward the left wall of the channel. Along y/d — 0.17, the depth decreases less than hm\x\/d ~ 0.1 and gradually increases to 0.1, i.e., overshooting the reduction, while along y/d = 1.30, near the left wall, the reduction is gradual without the overshoot. The flow patterns of the gravity current and internal bore in the horizontal plane at z/d = 0.27 are visualized a.nd presented in Figs. 17 and 18 respectively.
Descriptive
0.6
Hydrodynamics
of Lock-Exchange
Flows
229
the tip of thei partition wall (x=0)\
*^
0.3
0
O • © + ffl
(y/d) 1.30 0.73 0.17 -0.3 2-D (y/d=0.73)
0
£i
JC
Jbl C&-
Fig. 16. Variations of the depressed water depth behind the leading head of the gravity current. The depth /i m i„, as identified in Fig. 15, is normalized with the total flow depth d, which is plotted against the location of the leading head.
f=1.2s
f = 11.8 s
f = 7.1s
t= 18.9 s
Fig. 17. Time sequences of the flow patterns of gravity current (Fig. 15(a), R = 1700, F = 0.73) in the horizontal plane (the X-AJ plane) at the height z/d = 0.27. The location of the end of the partition wall (x = Q,y = 0) is marked by • . Note that the time origin (t — 0 s) is set at when the leading edge of the current passes at the position x = 0 (the end of the false partition wall).
230
H. Yeh & K. Wada
( = 0.0s
t~ 13.7 s
A
r = 8.0s
t- 16.1s
Fig. 18. Time sequences of the flow patterns of internal bore (Fig. 15(b), R = 1800, F = 0.78) in the horizontal plane (the x-y plane) at the height z/d = 0.27. The location of the end of the partition wall (x — 0,y = 0) is marked by • . Note that the time origin (! = 0 s) is set at when the leading edge of the current passes at the position x — 0 (the end of the false partition wall).
In Fig. 17, the gravity current shows the complex leading front pattern just as in Fig. 6. On the other hand, the leading front of the internal bore is smooth as shown in Fig. 18 (also see Fig. 12). Time sequences of Figs. 17 and 18 clearly show that once the front leaves the confined channel, the fresh water inflow is introduced from the region y < 0 into the channel. While the current is advancing within the confined channel (x < 0), the fresh water and salt water are simply exchanged within the confined channel just like the flow in a two-dimensional channel as shown in Fig. 14. Once the front escaped from the confined channel (x > 0), the fresh water intrudes from the side (y < 0) around the end of the right wall to supplement the volume loss due to the saline current outflow into open environment. This fresh water intrusion from the side around the end wall causes the sudden flow-depth reduction behind the leading head as well as inducing the strong three-dimensional mixing and upwelling near the end wall as shown in Fig. 15. For the internal bore, the roll-up (wave breaking) behind the head manifests itself as a white streak line in Fig. 18. Observation of the time evolution in Fig. 18 suggests that the roll-up is not aligned to the direction of the front, but the direction is oblique to the spreading current. Recall that the roll-up and the front are usually parallel in a two-dimensional channel as shown in Fig. 12.
236
H. Yeh & K. Wada
90 •
d
-o
G J---0 Q----H-Q
75
A
1 G.C.
o
3
60
•'": . t
45 \
\l.B.(*„ = 30 30
i
, 40
50
I.BJ0\=lcm)
O.Scm) ,
60
i
70
?(S)
80
Fig. 25. Temporal variations of the side interface slope f3 as indicated in Figs. 23 and 24; • , gravity current; D, internal bore with ho = 0.5 cm; o, internal bore with ho = 1.0 cm. The peak locations of dominant oscillations are marked by: — • — for the gravity current; D for internal bore with ho = 0.5 cm o for internal bore with ho = 1.0 cm.
The time variations of the side slope (3 at x = 2 cm as shown in Figs. 23 and 24 are plotted in Fig. 25. Note that the side slope (3 is much steeper, (3 > 45°, than the front slope as shown in Fig. 14. Also noted that, unlike the front slope of the gravity current being much steeper than that of the internal bore, the values of /3 for gravity current and internal bore are comparable. It is partly because the location of the measurements is right after the end wall (x = 2 cm) and its outflow momentum disallows the flow to spread at this point. The slope (3 for both cases fluctuates at comparable rates, the frequency 0.1 ~ 0.15 Hz. This pulsating side slope might be related to the eddy formation due to baroclinic torque on the side face. The eddy formation process is not continuous but discrete, i.e., the process repeats its creation and advection. This creation-advection cycle may have caused the distinct fluctuation of the side slope. Using Eq. (5) and based on the flow-visualization results described herein, the following flow patterns can be inferred for the gravity current and internal bore when they spread out from the confined channel. Once the current exits the channel into open environment, the effects evidently first appear near y « 0. Near the exit point x w 0, the side interface is nearly vertical and the fluid there appears to be vortical due to advection of fluid from the boundary layer along the tank's side wall (Figs. 23 and 24): The initial vorticity at the
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
237
Fig. 26. Generation and transport of fluid rotation near the exit to open environment. While the vortical flow along the side-wall boundary is advected to the open environment, the creation of vertical salt and fresh water interface at the exit induces fluid rotation in the negative x-direction due to baroclinic torque. The created fluid rotation is stretched and bent by the fresh-water inflow to form a vortex tube, which influences the roll-up pattern of gravity current and internal bore as shown in Figs. 17 and 18.
sidewall must be in the negative z-direction. At the same time, once the current exits from the channel, the vertical side interface induces strong baroclinic torque l/p 2 Vp x Vp, which creates the vorticity in the negative ^-direction and causes slump of the interface in the negative y-direction and induction of fresh-water intrusion in the positive y-direction. Such fluid-rotation patterns near the exit (x = 0 and y = 0) is depicted in Fig. 26. The induced freshwater intrusion from the right provides the dominant supply into the channel to compensate the saline loss. Hence, the flow pattern near the exit becomes three-dimensional. At the same time, the created vortical flow at the side interface due to baroclinic torque is advected with this fresh-water intrusion,
238
H. Yeh & K. Wada
and then vortex tubes are being bent and stretched. This fresh-water advection adjacent to the saline interface is initially in the positive ^-direction due to the initial saline-flow momentum, then is bent inward (in the y-direction) and accelerated due to the flow convergence into the channel. Hence, the initial fluid rotation in the negative ^-direction is bent and stretched toward the negative y-direction as depicted in Fig. 26. This bending causes the roll-up behind the internal bores shown in Figs. 18 and 22, and also the rotation causes the fresh-water down welling, resulting the substantial reduction of the saline depth behind the head (see Figs. 15 and 16). This can explain why the roll-up and the depth reduction propagate toward the y-direction, because the roll-up is caused by the fresh-water intrusion from the right-hand side, instead of the roll-up from the leading head (Note that in the two-dimensional channel, the roll-up is caused by the accumulation of baroclinic torque generated along the front face). Inside of the channel near the right wall, we observed strong mixing and upwelling (see Figs. 15, 19, and 20). This upwelling must be due to the
Fig. 27. Schematic flow pattern near the exit. The upwelling of the mixed fluid occurs by the secondary current generation due to the fresh-water separation eddy formation (see Figs. 15, 19, and 20).
Descriptive
Hydrodynamics
of Lock-Exchange
Flows
239
formation of the secondary current caused by flow separation of fresh-water inflow around the tip of the right wall. The inflow velocity is fast near the surface and it decreases in depth: The inflow velocity must diminish just above the saline outflow. Hence, just as the separation flow pattern in an openchannel flow (Yeh, et al., 1988), the separation eddy is larger near the surface and smaller near the fresh-water-and-saline interface as depicted in Fig. 27. The flow divergence of the upper fresh water causes the upwelling of the lower fluid. Nonetheless, the mixed fluid that appears in the upwelling must be originated at the side interface near the exiting point. Note that active mixing near i/fsO can be seen in Figs. 23 and 24: The mixing is due presumably to baroclinic torque as well as the advection of vortical fluid from the side wall as depicted in Fig. 26. Such mixed fluid may be transported by the fresh-water inflow to the area of the upwelling. 7. Conclusion It is reconfirmed that the gravity current near the leading head is significantly different from that of the internal bore. When the gravity current or internal bore spreads from the two-dimensional channel into open environment, the flow pattern of each case is similar except at the leading front. The leading front of the gravity current is complex, while that of the internal bore is smooth. The complex flow pattern associated with the spreading appears to be influenced significantly by the upper fresh-water inflow to the channel and by the fluid rotation generated by baroclinic torque at the flow exit. The vigorous mixing and upwelling near the flow exit within the confined channel must be caused by the fresh-water inflow. Acknowledgment The photographs presented herein were taken at the University of Washington by C. Grandinetti, S. Lingel and K. Wada. The support from the Washington Sea Grant Program is acknowledged. References Benjamin, T. B. (1968). Gravity currents and related phenomena. J. Fluid Mech. 31: 209-248. Bernal, L. P. (1981). The coherent structure of turbulent mixing layers. I. Similarity of the primary vortex structure; II. Secondary streamwise vortex structure.
240
H. Yeh & K. Wada
Report, Graduate Aeronautical Laboratories, California Institute of Technology, Pasadena, Calif, pp. 92. Breidenthal, R. (1981). Structure in turbulent mixing layers and wakes using chemical reaction. J. Fluid Mech. 189: 1-24. Chu, V. H. and R. E. Baddour (1977). Surges, waves and mixing in two-layer density stratified flow. Proc. the 17th IAHR Congr.: 303-310. Denton, R. A. (1990). Accounting for density front energy losses. J. Hydr. Engr. 116: 270-275. Fischer, H. B., E. J. List, R. C. Y. Koh, J. Imberger and N. H. Brooks. (1979) Mixing in Inland and Coastal Waters. Academic Press, New York. Grandinetti, C. (1992). Gravity current and internal bores. M. S. thesis, Univ. of Washington, Seattle, pp. 82. Jirka, G. H. and M. Arita (1987). Density currents or density wedge: boundary-layer influence and control methods. J. Fluid Mech. 177: 187-206. Klemp, J. B., R. Rotunno, and W. C. Skamarock (1994). On the dynamics of gravity currents in a channel. J. Fluid Mech. 269: 169-198. Lamb, H. (1932). Hydrodynamics, Cambridge University Press, Cambridge. 6th ed. Lasheras, J. C , J. S. Cho, and Y. Maxworthy (1986). On the origin and evolution of streamwise vortical structures in a plane, free shear layer. J. Fluid Mech. 172: 231-258. Lighthill, J. (1980). An Informal Introduction to Theoretical Fluid Mechanics, Clarendon Press, Oxford. Lin, S. J. and G. M. Corcos (1983). The mixing layer: deterministic models of a turbulent flow. Part 3. The effect of plane strain on the dynamics of streamwise vortices. J. Fluid Mech. 141: 139-178. Maxworthy, T. and S. Monismith (1988). Differential mixing in a stratified fluid. J. Fluid Mech. 189: 571-598. Serrin, J. (1959). Mathematical principles of classical fluid mechanics. In Handbuch der Physik VIII/1, Springer-Verlag, Berlin, pp. 125-263. Simpson, J. E. (1987). Gravity Currents: In the Environment and the Laboratory. Ellis Horwood Ltd, Chichester. Wilkinson, D. L. and I. R. Wood (1972). Some observations on the motion of the head of a density current. J. Hydraul. Res. 10(3): 305-324. Wood, I. R. and J. E. Simpson (1984). Jumps in layered miscible fluids. J. Fluid Mech. 140: 329-342. Yeh, H. (1991). Vorticity-generation mechanisms in bores. Proc. Roy. Soc, Lond. A 432: 215-231. Yeh, H. (1995). Free-surface dynamics. In Advances in Coastal and Ocean Engineering, ed. P. L.-F. Liu, Vol. 1, World Scientific, Singapore, pp. 1-75. Yeh, H. H., W. Chu and O. Dahlberg (1988). Numerical modeling of separation eddies in shallow water. Water Resources Research, 24(4): 607-614. Yeh, H. H. and K.-M. Mok (1990). On turbulence in bores. Phys. Fluids A 2: 821-828.