Open access
Technical Papers
Apr 1, 2016

Large Eddy Simulation of Unidirectional and Wave Flows through Vegetation

Publication: Journal of Engineering Mechanics
Volume 142, Issue 8

Abstract

Massively parallel large eddy simulation (LES) experiments were conducted to study the flow fields developed by unidirectional flow over submerged vegetation and wave flow-through emergent vegetation. For the submerged vegetation, vertical profiles of mean and turbulent horizontal and vertical velocities were found to be in good agreement with laboratory experiments. Canopy-averaged bulk drag coefficient calculated from the depth-integrated forces on the cylinders compares well with empirical measurements. For the emergent vegetation, wave-induced drag forces were calculated, and the inertia and pressure-drag coefficients were compared with laboratory experiments, which were found to be in good agreement. Vertical variation of forces and moments about the stem base are presented and compared with a single stem case under a variety of wave conditions and Keulegan-Carpenter (KC) numbers. It is seen that at low KC numbers the effect of the inertia force is significant. The vertical variation of the velocity field near the free surface, within two to three diameters from the cylinder center, was found to be considerably influenced by the free surface flux between the wave trough and the crest, over a time period, which creates a predominant recirculation zone behind the cylinder in the direction of wave propagation.

Introduction

Flows through riverine and coastal vegetation have been areas of active research among civil engineers for the last half century. Submerged riverine vegetation exists largely in unidirectional flow environments and induces vegetation drag, reducing the mean flow velocities within the canopy with respect to the free upper layer, whereas emergent vegetation canopies in coastal wetlands are most effective in attenuating the wave height by dissipating the wave energy. In both cases the forces exerted by the plant stems on the surrounding flow strongly influence the mean and instantaneous flow field, species, and sediment transport processes. Experimental work in unidirectional flow over vegetation (Kouwen et al. 1969; Tsujimoto et al. 1991; Dunn et al. 1996; Fairbanks and Diplas 1998) has successfully established that the mean horizontal velocity profile within a submerged vegetated layer shows significant departure from the universal logarithmic law, and the flow over is more akin to a mixing layer flow analogous to what is seen in atmospheric canopies rather than a perturbed boundary layer flow (Ghisalberti and Nepf 2002, 2004, 2006; Nepf and Ghisalberti 2009). It was also found that, apart from the variation in the mean quantities, large spatial variation in the vertical profiles of kinetic energy and Reynolds stresses exist within the canopy (López and García 2001).
Experimental studies looking into wave attenuation through emergent and submerged artificial as well as live vegetation (Asano et al. 1988; Kortlever 1994; Lowe et al. 2005a, b; Augustin et al. 2009; Chakrabarti and Smith 2011; Jadhav and Chen 2013; Jadhav et al. 2013) found that the wave height decay and resulting turbulence is closely related to population, diameter of the stems, degree of submergence, rigidity of the stems, and wave conditions. Recently Hu et al. (2014) showed that the ratio of the imposed current velocity to the amplitude of the orbital velocity determines whether a following current increases or decreases the vegetation drag. They used Morrison’s equation to calculate a period and spatially averaged drag coefficient from known force data—the first observation of its kind.
The first analytical model based on the wave period–averaged energy conservation was conceptualized by Dalrymple et al. (1984) in which the local flow velocity was assumed to follow linear wave theory. The drag force was represented by the quadratic law using the wave orbital velocity, whereas the inertia component was assumed to vanish when averaged over a wave period. Subsequent improvement in modeling efforts (Kobayashi et al. 1993; Mendez and Losada 1999, 2002; Maza et al. 2013; Chen and Zhao 2012; Ma et al. 2013; Wu et al. 2013; Zhao and Chen 2014; Marsooli and Wu 2014; Zhu and Chen 2015) introduces a drag force–based sink term in the momentum equation to represent vegetation drag. Although this class of models is relatively easier to implement in the governing equations and is able to give reasonable results on mean flow quantities, it has several limitations. First, because the drag coefficient is unknown in the governing equations, it requires calibration against known wave height decay curves from wave flume experiments, and the calibrated coefficients thus become a function of the fidelity of the correlation of the modeled and experimental data. Second, the momentum loss from turbulent interaction of the flow between the stems and bed friction are all lumped into a drag force term, which may be one of the reasons for the large scatter commonly seen between studies of different researchers. Last, these models do not give any insights into the nature of eddy shedding from vegetation stems and resultant turbulent flow structure within canopies.
In view of the limitations of the previously mentioned models, a more direct approach to solving the wave vegetation interaction problem is proposed in the form of high-resolution large eddy simulation (LES) (Sagaut 1998)-based turbulence modeling using a three-dimensional (3D) Navier-Stokes equation in which the flow structure within the canopy is resolved up to a reasonable order of accuracy verified by first-order (mean) and second-order (turbulent quantities) from known experiments. The only assumption in this model is that the large eddy flow features, which develop around the cylinders, are the principal energy extractors from the overall flow. This is the first study simulating and investigating free surface nonlinear wave flow around an array of rigid vertical cylinders by resolving the flow field in the canopy. The drag force is calculated by integrating the time-varying instantaneous pressure field around each cylinder in the vegetation array, and no calibration is required.
The LES is an intermediate approach between Reynolds-averaged Navier-Stokes (RANS) or unsteady RANS (URANS) and direct numerical simulation (DNS). RANS-based models (Shimizu and Tsujimoto 1994; Naot et al. 1996; López and García 1997; Fischer et al. 2001; Neary 2003) used earlier required explicit specification of momentum sinks to account for the drag force. On the other hand, DNS is still restricted to small-scale domains [O(103) to O(102)m] and low Reynolds number environments (Liu et al. 2011, 2015). The LES method has been applied before to study the unidirectional flow around arrays of emergent and submerged vegetation stems (Cui and Neary 2002; Stoesser et al. 2009; Palau-Salavador et al. 2010). These LES studies offered a higher degree of detail into the flow and an opportunity to answer hitherto unanswered research questions, albeit at the cost of greater computational resources, when compared with URANS-based models. However, LES studies in vegetated flows are still limited to relatively moderate Reynolds numbers [O(102103)] because of the need for very fine grids near the vegetation elements and corresponding very small time steps as well as the sheer size of the problem. Luckily, the waves in the estuarine zone in which vegetated coastal wetlands exist exhibit an orbital velocity Reynolds number in that range. The LES studies of this nature can only be run at present in dedicated high-performance computing (HPC) clusters with large numbers of nodes, fast networks, and large storage. The simulations in this paper were run in a 146 TFlops peak performance cluster with 440 compute nodes, with each node having two 8-Core Sandy Bridge Xeon 64-bit processors (Intel, Santa Clara, California) operating at 2.6 GHz with 32 GB of RAM and sharing a 40Gb/s Infiniband (Mellanox Technologies, Sunnyvale, California) network interface between them. Typically 64–160 nodes are allocated for each run based on scalability curves and depending on the problem size.
In this paper the principal aim is to validate the 3D LES model based on OpenFOAM (Weller and Tabor 1998) when applied to vegetated flow and to apply the model to study the wave interaction with an array of cylinders with an emphasis on understanding how the drag force on the cylinder varies vertically and the influence of the free surface hydrodynamics in modifying the velocities around the cylinder. The first part of the paper deals with the comparison of mean and turbulent velocities for the unidirectional flow over an array of submerged cylinders against the experimental results of Liu et al. (2008). The distribution of depth-integrated drag force in the canopy has been investigated, and the bulk drag coefficient calculated from the modeled forces has been compared with the empirically derived relations by Ghisalberti and Nepf (2004), relating the submerged vegetation canopy drag coefficient to that in emergent vegetation. The agreement was found to be within 5% and further validates the empirical equation. The second part shows the application of the model to study the wave interaction with an array of cylinders and a single cylinder. The drag and inertial coefficients induced by the wave forces acting on a cylinder are obtained by fitting Morrison’s equation to the LES model computed force data. These are then compared with those obtained in the experiments of Hayashi and Chaplin (2012). Insights into the vertical variation of the wave forces, vertical variation of the amplitude of wave velocity, and the complex vortex shedding mechanism from a single cylinder are presented for a variety of wave heights and time periods. An attempt has been made to explain the complex hydrodynamics at the free surface caused by nonlinear wave interaction with a cylinder, either isolated or in an array. This work is expected to be the beginning of several future studies looking at wave and vegetation interaction using LES. With the development of HPC infrastructure this approach has the advantage of revealing several important canopy-scale phenomena, such as the role of the spacing between stems, characteristic vortex features in canopy associated with wave conditions, relative importance of the orbital components in the energy reduction, and the unique flow features caused by wave current-vegetation interactions.

Numerical Methods

All the experiments in this paper were conducted using OpenFOAM v. 2.1.1 (Weller and Tabor 1998). For the unidirectional flow experiments over submerged vegetation, the 3D incompressible LES-filtered continuity and the Navier-Stokes equation were used to resolve the flow field. The equations can be expressed as
·u¯=0
(1)
u¯t+·(uu¯)=1ρp¯+·ν(u¯+u¯T)+·τsgs
(2)
where vector notations are represented as bold-faced characters with u=(u,v,w); and u, v, and w are the magnitude of velocities in the X-, Y-, and Z-directions, respectively. Also, t = time variable; ρ = density of water; p = pressure; ν = kinematic viscosity; and τsgs is the subgrid scale–modeled shear stress. The LES spatial filter operator, written generically is
ϕ¯(x)=Ωϕ(x)G(x,x)dx
(3)
where Ω = fluid domain; and G = filter function that specifies the scales of the resolved eddies. In OpenFOAM, LES filtering is provided implicitly by the finite-volume discretization itself and the filtering operator in Eqs. (1) and (2), represented by the overbar, becomes
ϕ¯(x)=1VVϕ(x)dx
(4)
for all (x ϵ V), V is the volume of a computational cell, and the filter function is the simple box-hat filter
G(x,x)={1VifxϵV0otherwise
(5)
For the free surface wave flow experiments over emergent vegetation, in addition to solving Eq. (1), the LES-filtered combined equation for the air and water flows are solved
ρu¯t+·(ρuu¯)=p*¯g·xρ+·[μu¯+ρτsgs]
(6)
In Eqs. (2) and (3) the subgrid-scale stress term (τsgs) was modeled by the dynamic Smagorinsky model (Germano et al. 1991) with local averaging of the Smagorinsky constant. The effective viscosity was clipped such that it is never allowed to be less than zero to keep backscatter effects resulting from negative subgrid-scale viscosities physically consistent. This is the same LES model that was also used by Stoesser et al. (2009), and their results have been presented later as a comparison.
In the volume of fluid (VOF) (Hirt and Nichols 1981) method, because air and water are treated in a single-phase calculation throughout the domain, the physical properties of the mixture [ρ and μ in Eq. (3)] are calculated as the weighted average of their respective properties with the weighting done by the phase fraction
ρ=αρw+(1α)ρa
(7)
μeff=αμeff,w+(1α)μeff,a
(8)
where the subscripts represent water (w) or air phase (a) values; and the effective viscosity is μeff=μ+μsgs. The phase fraction (α) can have values within [0,1] with α=0 signifying fully air phase and α=1 signifying fully water phase with cells having intermediate values representative of partially air- and water-filled zones. OpenFOAM uses a modified version of the two-fluid Eulerian model approach typically used in two-phase flows to describe the transport of the phase fraction. In the multidimensional universal limiter for explicit solution (MULES) (Berberović et al. 2009), the transport of the phase fraction can be represented by a single combined evolution equation, which is written as
αt+·(αu¯)+·[u¯rα(1α)]=0
(9)
where u¯=αu¯w+(1α)u¯a and u¯r=u¯wu¯a represent the relative velocity vector termed compression velocity, which in turn introduces an additional convective term referred to as the compression term to the governing equation. This term can be properly tuned to achieve a very high interface resolution avoiding the need for special interface treatments used in a classical VOF method and makes this VOF technique very effective and unique, although it requires significantly small time steps [O(103104)s] for practical wave flow problems.
The numerical domain was discretized using the finite-volume method with both spatial and temporal discretizations second-order accurate. The pressure implicit with splitting of operators (PISO) (Ferziger and Peric 2001) algorithm with two pressure corrector steps and two nonorthogonal correctors was used for the pressure-velocity solver. For the solution of the linear algebraic equations, the preconditioned conjugate gradient method (PCG) with diagonal incomplete Cholesky (DIC) (for symmetric matrices) and preconditioned biconjugate gradient (PBiCG) with diagonal incomplete lower upper (LU) factorization (for asymmetric matrices) were used as the preconditioners.

Numerical Domain and Boundary Conditions

For the uniform flow over submerged vegetation experiment, the domain size was similar to the experiments of Liu et al. (2008) and numerical simulations of Stoesser et al. (2009). Length, width, and height of the domain were, respectively, 40, 20, and 18D, in which D=6.35mm is the diameter of each cylinder. Sixteen cylinders were arranged in a staggered pattern with the distance between them being 10D [Figs. 1(a and b)]. For the results presented in this paper, measurements were taken at the six locations marked 1–6, as in Fig. 1(b). At the inlet, outlet, and sides of the domain the cyclic boundary condition for velocity and pressure were imposed, whereas the top was given a full-slip boundary condition, treating it as a rigid lid. A no-slip boundary condition was imposed for the bottom of the domain and on the cylinder walls. To correctly simulate the near-wall turbulence, Spalding’s smooth wall function (Spalding 1961) was used to calculate the near-wall turbulent eddy viscosity. Flow was driven by specifying the pressure gradient at every time step, such that the bulk velocity in the domain (Ub) was kept constant. The domain had a total of 12,764,590 finite-volume cells with the near-wall resolution of 0.45 mm corresponding to y+=46. The mesh was refined in the X, Y- plane near the cylinder walls [Fig. 2(a)], whereas in the Z-direction refinement was done to capture the bottom boundary layer using a geometric expansion. The near-wall aspect ratio was maintained at ΔxΔyΔz=111. The channel, cylinder, and canopy Reynolds numbers were, respectively, Rchannel=38,000, RD=2,100, and Rcanopy=1,340.
Fig. 1. Numerical domain and sampling locations: (a) uni-directional flow domain; (b) sampling locations: D is the cylinder diameter; (c) free surface wave flow domain; (d) sampling locations: S=10D points 2 and 3 are at 1.5D distance from the center
Fig. 2. Horizontal and vertical mesh sections: (a) mesh section at half of cylinder height within the array; (b) vertical section through flume center for free surface wave flow experiments showing refinement near the bottom boundary layer and near the free surface
For the wave flow experiments over emergent vegetation, the waveFOAM toolbox (Jacobsen et al. 2012) was used. The domain was 0.25 m wide, 0.75 m high, and (1.8+3L)m long, where L is the wavelength of the incident wave, with the central 1.8 m in the flume occupying the vegetation zone as in the original laboratory experiment [Fig. 1(c)]. The free lengths (before and after the vegetation zone) are considerably shorter than in the original experiment to reduce computational cost. Also, because one is mostly concerned with the hydrodynamics inside the vegetation zone, it is not important to have such a long domain. The width of the domain is 0.25 versus 0.8 m in the original experiment, because sensitivity analysis with a larger width did not show any appreciable differences to the force and velocity measurements at the central cylinder. Fifty-three cylinders were placed in a staggered arrangement, such that the nearest neighbor distance between them was 10D [Fig. 1(d)]. Velocity measurements were taken from locations marked by X, as in Fig. 1(d), and averaged to yield a representative velocity for the evaluation of drag coefficients from force measurements of the central cylinder (53). Surface elevation readings taken from these locations, on the same X-coordinate as the central cylinder on either side, were also averaged to represent surface elevation reading at the cylinder location. For the isolated cylinder experiments only cylinder 53 was placed in the domain. Vertical profile of 3D velocity data from Locations 1–4 in Fig. 1(d) have been used to compare maximum wave velocities in the canopy and the effect of the cylinders on the wave flow. At the inlet the surface elevation of waves was set by specifying the phase fraction (α) using Stokes’ second-order wave theory. To damp the waves generated in the numerical flume a numerical sponge layer or relaxation zone is needed. Relaxation zones, for both isolated cylinder and array cases, extended up to one wavelength into the domain from the inlet and outlet. For the relaxation zone a function similar to that used by Engsig-Karup (2006) and shown in Eq. (10) was used, where exponent p was chosen as 3.5 based on sensitivity analysis of the reflection coefficients with w and σ as the weighting function and the local distance function, respectively (Jacobsen et al. 2012)
w=1σp
(10)
The local Courant number correction method of Seng et al. (2012) was applied as an additional temporal correction within the relaxation zones. The mesh was refined in the X, Y-plane similar to Fig. 2(a) and also in the Z-direction near the bottom boundary layer as well as near the fluid cylinder interface [Fig. 2(b)]. Minimum cell sizes outside the central 2.0-m zone were double the size of the largest cell within the vegetation zone and were expanded near the inlet and outlet boundaries. Local refinement was done at the still-water level (SWL) so that the maximum oscillation of the free surface remained within the refined zone. The aspect ratios of cells at the free surface were kept as 111, except within the relaxation zones in which the cells were expanded in the horizontal direction.
The sides of the domain had full-slip walls, whereas the top was given an atmospheric pressure condition. The bottom of the domain and cylinder walls had the no-slip condition along with Spalding’s wall function (Spalding 1961) for the turbulent viscosity term. The wave conditions and mesh resolutions are given in Table 1. Linear wave theory was used to calculate the maximum velocity at free surface, which is used to calculate the nondimensional parameters in the table. The near-wall aspect ratio was maintained at ΔxΔyΔz=112.5. Sensitivity analysis of the near-wall mesh requirements indicated that a minimum of 64 points along the cylinder circumference was required to accurately simulate the wake structures and account for the pressure fluctuations near each cylinder. A near-wall resolution of D/20=0.5mm was maintained along the circumference with the minimum cell size in the wall-normal direction being D/50=0.2mm, and the cells were stretched away from the wall. For both the unidirectional and wave flow simulations, time stepping was done adaptively, such that the maximum Courant number [max(Cr)] in the domain does not exceed 0.2. The Courant number (Cr) for a given cell with volume V is defined as Cr=[Σi|Ui·Ai|/0.5V]Δt, where the velocity and area vector of the ith face are Ui and Ai. The factor 0.5 ensures that the mean of the velocities is taken for opposite faces when calculating the sum. The numerical time-stepping interval (Δt) varied between 106 and 105s for the unidirectional flow runs and between 106 and 0.5×103s for the wave flow runs.
Table 1. Unidirectional Flow: Force Distribution among Cylinders in the Channel
Cylinder numberFpi¯×103NFvi¯×103NFpi,rms×103NFvi,rms×103N
116.01770.61983.10820.0861
216.26770.64743.55700.0996
316.11020.64432.66620.0760
416.16430.64672.97320.0855
516.21030.62603.16230.0864
616.16430.64543.41710.0957
716.17510.64522.73680.0770
816.19240.64832.92100.0835
916.13650.62233.16260.0846
1016.05570.64023.48490.0970
1116.01230.64182.72790.0760
1216.16430.64543.41710.0957
1316.03320.61953.15890.0853
1416.30500.64953.57880.0989
1516.01550.64252.70230.0779
1616.06960.64362.92190.0838

Results

The goal of using 3D LES modeling is to better understand the hydrodynamics within vegetation canopies and quantify the variation of forces acting on the stems. However, because of the scarcity of available laboratory experiments in the literature, which simultaneously document mean and turbulence quantities as well as drag forces, it was decided to conduct separate validations for unidirectional flow in which the mean and turbulence quantities could be verified and free surface wave flows in which the force coefficients could be validated.

Unidirectional Flow

The experimental conditions were chosen similar to the laboratory experiments of Liu et al. (2008), where the bulk velocity (Ub) in the channel was specified as 0.33m/s. An initial ramp up time of 18 flow-through periods (tf=l/Ub) were allowed before data collection started. The choice of this period was based on the fact that the mean and turbulence velocity profiles were devoid of any initial effects after 18 flow-through cycles. The mean velocity results became independent of the averaging period after 45 flow-through cycles beyond the initial ramp time, whereas the mean turbulence quantities became invariant of the averaging period after 50 flow-through periods beyond the initial ramp time. The independence of the results on the averaging period was quantified by comparing the difference in the RMS deviation (RMSD) with respect to the experimental data for the vertical profiles at the six chosen locations for each incremental averaging period with a step size of 5. When the difference in the RMSD between two profiles obtained from two successive averaging periods became less than 1%, the latest averaging period was chosen as being sufficient for data analysis purposes. Thus, the data presented in this paper were obtained by averaging over 55 flow-through periods after the initial 18 flow-through periods for flow ramping. Temporal data were collected at 333 Hz, which was sufficient to trap the tail of the Kolmogorov 5/3 spectrum.
In Fig. 3 the first and third row of panels from the top show the vertical profile of the variation of mean horizontal (U) and vertical (W) velocities, nondimensionalized by the bulk velocity (Ub) at each of the six locations shown previously in Fig. 1(b). The second and fourth rows show the vertical profiles of the turbulence component of the velocities in the two directions nondimensionalized by the shear stress defined as U*=ghS, where h=0.114m was the height of the domain as well as the depth of water; and S=0.003 was the slope of the channel in the original laboratory experiment. The turbulent velocities were calculated as RMS of the observed velocities
Ut=1ni=1n(UiU)
(11a)
Wt=1ni=1n(WiW)
(11b)
where Ui and Wi = resolved instantaneous horizontal and vertical velocity values.
Fig. 3. Unidirectional flow validation: comparison of nondimensionalized mean horizontal velocity (U), turbulent horizontal velocity (Ut), mean vertical velocity (W), and turbulent vertical velocity (Wt) for the six locations (V1, V2, V3, V4, V5, and V6) as marked in Fig. 1(b) with the experimental data adapted from Liu et al. (2008) and the high-resolution LES simulation of Stoesser et al. (2009); the black horizontal line signifies the top of the vegetation canopy; Ub and U* are, respectively, the bulk and shear velocities in the channel
Generally, it is observed that the mean horizontal velocities at all six locations match well with the experimental data and also compare favorably with the numerical results of Stoesser et al. (2009), considering the fact that the latter used a mesh with about twice the resolution as the present simulation. The streamwise velocities within the canopy layer are considerably lower than those in the fluid layer above the canopy and follow the general trend observed in flow over submerged vegetated canopies. The fluid on the top of the canopy accelerates and produces the classical hyperbolic tangent profile with the inflection point close to the top of the canopy. All six positions show the same global hyperbolic type profile, which suggests the spanwise homogeneity of this phenomenon, with the locations just behind (V1) and before (V2) the stem showing large gradients of horizontal velocity near the top of the canopy. This indicates the effect of the recirculation zone, which can be defined as the sheltered region within the wake of a cylinder, in the dominant flow direction, within which the centerline mean streamwise velocities are negative. It is worthwhile to note that the mean horizontal velocity within the recirculation zone is actually better predicted in this model than in Stoesser et al. (2009) who have noted that the accuracy of prediction of this region is strongly dependent on the choice of the subgrid-scale (SGS) model, LES data averaging period, or measurement period in the original experiment. In initial sensitivity runs it was found that the accurate prediction of the mean and turbulent quantities in the recirculation zone are much more sensitive to the near-wall mesh than the averaging period or the choice of the SGS model, assuming of course the averaging period is large enough so statistically steady-state turbulence profiles have been reached within a 1–2% error. Hence, the model consistently underpredicts the gradients of mean and turbulence quantities in the velocity profiles within the recirculation zone and even more so at vertical V1. An important goal of this research is to use LES simulations, with mesh sizes and temporal resolutions, which can successfully simulate experiments conducted in typical wave flumes found in engineering laboratories, within a reasonable level of accuracy. These experiments may either be in field scales or prototype scales of larger coastal wetland canopies mimicking the height, density, and diameter of the rigid portion of the stems. As a secondary condition, the total size of the problem in these simulations should be such that they can be run in conventional HPC clusters with relatively moderate computational effort. Hence, the near-wall resolution, which yields acceptable results for mean quantities in the majority of the locations, is considered satisfactory.
The streamwise turbulence velocity profiles (second row from the top in Fig. 3) agree very well with the experimental data, except at location V1, because of the previously explained reason. The peak of the intensities occurs at or near the stem tops with the peak lying a little below the tops at the free locations (V2 and V5), suggesting that streamwise vortices advected from preceding stems are pushed down when they reach the sideways free locations. It is also noted that the LES model slightly underpredicts the peaks at those two locations as was observed by Stoesser et al. (2009). The present model predictions are found to be better than those by Stoesser et al. (2009) in trapping the vertical gradients of the turbulent horizontal velocities at the free locations (V2 and V5). Also, in the above canopy region the turbulence streamwise velocities are better predicted consistently, particularly near the full-slip wall boundary at the top. The difference is caused by the full-slip boundary condition set at the top versus the symmetry condition as imposed by Stoesser et al. (2009), who instead explained the overprediction of turbulence intensities near the rigid lid to coarse resolution and to large, roughness scale turbulence elements originating at the actual free surface, which is treated as a rigid lid for the experiment presented in this paper. Simulations run with symmetry condition in this case produced similar overshoots caused by reflection of turbulent vortices back into the domain, and it is recommended that a full-slip top wall is a much better boundary condition than symmetry for setups of this type.
The vertical profiles for the vertical velocity (third row from top in Fig. 3) agree reasonably well with the experiment except within the recirculation regions (V1 and V6) in which the model does not predict the peaks very well. The vertical velocities also seem to be a little overpredicted both within and above the canopy. The LES model captures the peaks of the profile very well for the sheltered zones just outside the recirculation region (V3 and V4), but it seems to miss those in the free zones (V2 and V5) and is attributed to lower resolution at these locations.
Velocity profiles for the turbulent component of the vertical velocity (fourth row from top Fig. 3) are found to agree well consistently for all the locations, except for the recirculation zones (V1 and V6). The underprediction of the turbulent quantities within the recirculation zones is caused by a grid resolution lower than that used by Stoesser et al. (2009). The larger cell sizes within this zone are unable to trap the eddies with the peak turbulent energy, resulting in an under-prediction of the turbulent velocities. The peak of the profiles occur just below the canopy top for the free zones (V2 and V5) and at the canopy top for the sheltered zones (V1, V3, V4, and V6) and compares well with the results of Stoesser et al. (2009).
In addition to sampling velocity profiles, the depth-integrated forces on each cylinder were computed as a sum of pressure (Fp) and viscous (Fv) forces as Ftot=Fp+Fv
Fp=hη02πpcos(ϕ)rdϕdz
(12a)
Fv=hη02πτeffsin(ϕ)rdϕdz
(12b)
where τeff=(τ+τsgs)dU/dn|cyl-wall is the effective shear stress on the cylinder wall with the gradient calculated with the wall parallel velocity. Here r and ϕ are the radius of the cylinder and the angle made by the infinitesimally small element located along the cylinder surface on which the forces act, with the axis of the cylinder in the X, Y-plane. The force data were sampled at 50 Hz and averaged over 55 flow-through periods to yield a representative steady-state force. The time-averaged and RMS forces for each of the cylinders is shown in Table 1. The mean forces showed little deviation from cylinder to cylinder with the percentage coefficient of variation (σ/μ×100), where σ and μ are, respectively, the standard deviation and mean of the forces on all cylinders, being about 0.6%, whereas the same for the RMS values was about 10%. The canopy-averaged mean force, Ftot¯=1/Ni=1N(Fpi¯+Fvi¯), was used along with the mean velocity of the cells immediately at the top of the canopy (Uhv¯) to compute the canopy representative drag coefficient (Tanino et al. 2005) given by Cd=Ftot¯/0.5ρUhv¯2Dhv, where hv=12D. The calculated values for Ftot¯, Uhv¯, and Cd from the model results were 16.77×103N, 43.2cm/s, and 0.371, respectively. This value of Cd has been compared with the empirically derived coefficient from Ghisalberti and Nepf (2004). The vertical variation of the ratio ζ=Cd(z)/CdA is given in Ghisalberti and Nepf (2004) as
ζ={1.4(zh)2.5+0.45,if0x0.764.8(zh)+4.8,if0.76<x1
(13)
where CdA=(1+10Rhv,D2/3)/1.16[1.169.31(ad)+38.6(ad)259.8(ad)3]=0.609 is the empirical expression for the bulk drag coefficient of a random, emergent array of cylinders as obtained by Nepf (1999), and Rhv,D=Uhv¯D/ν=2,743 is the Reynolds number using the velocity at the canopy top (Uhv¯). The value of ad for the present case is 0.071.
To derive the value of the depth-averaged drag coefficient, the expression in Eq. (8) is integrated as follows:
ζ¯=10.7600.76[1.4(zh)2.5+0.45]d(zh)+110.760.761[4.8(zh)+4.8]d(zh)=0.633
(14)
which gives a depth-averaged drag coefficient of Cd,Nepf=0.386. This was found to be in close agreement with the model predicted value of Cd=0.371, with a percentage difference of 3.9%. This analysis, in addition to providing an estimate of range of bulk drag forces, also validates numerically the method of Ghisalberti and Nepf (2004) for drag force estimation in submerged canopy flows and lends confidence to the present model, which can be used to supplement drag coefficient laboratory data sets of both submerged and emergent vegetation canopies in unidirectional flow in the future.

Free Surface Wave Flow

Six different wave conditions were simulated, as shown in Table 2, with each case run for an array of cylinders, an isolated cylinder, and a flume with no cylinders. The linear dimensions of the flume were the same (0.25 m wide, 0.75 m high, and 1.8+3Lm long, where L = wavelength of the particular case) for a given case for the three setups, with the same refinement regions (near the free surface and the bottom boundary) as described earlier. Table 2 also shows the problem sizes in terms of the number of cells for each experiment. The degree of wave nonlinearity is represented by the Stokes’ parameter (ϵ=kH/2) and Ursell number (Ur=H/h/(h/L)2=HL2/h3), where H = wave height at the wavemaker; h = uniform water depth; and L = wavelength calculated by the dispersion relationship. For a given h/L the degree of nonlinearity increases with an increase in Ur, whereas ϵ gives an absolute estimate of the degree of nonlinearity among different h/L cases, with higher values indicating greater nonlinearity. Velocity data were sampled at the four locations marked in Fig. 1(d) for both the array and isolated cylinder experiments. In addition to the isolated cylinder runs, velocity data were also sampled in the central vertical plane stretching from 25 to 25D with the cylinder placed at the origin. This was done to quantify the relative influence of the presence of the cylinder in modifying the wave field in either direction and compare it with the array cylinder velocity fields. Pressure (p) and velocities were collected for all the cells around the cylinders. Waves were started with a still-water condition in the flume, and it took approximately 18 wave periods from the start of the wavemaker to eliminate the effect of the initial conditions. The data were sampled at a frequency of 50/THz, where T is the wave period of the wave conditions simulated over a period of 35 complete wave periods after the initial spin-up time of 20 wave periods. The choice of the averaging period was made because about 28–32 wave cycles were required for averaging before the phase-averaged force coefficients for the central cylinder showed less than 1% variation for a subsequent increase in number of averaging cycles. Thus, a total of 55 waves were introduced into the domain, and averaging was done over the last 35 wave periods. Because the goal in this paper is to understand the hydrodynamics and resultant forces associated with the mean wave energy, the data were analyzed and frequencies larger than the second-order peak of the surface elevation spectrum were filtered out using a fast Fourier transform algorithm in MATLAB.
Table 2. Free Surface Wave Flow: Experimental Conditions and Problem Size
NumberH (cm)T (s)KCRβkH/2h/LUrArray (number of cells)Isolated (number of cells)Flume (number of cells)
13.00.969.1732800.0660.420.2814,399,2002,735,400203,500
26.00.9618.31,464800.1320.420.5614,399,2002,735,400203,500
37.50.9622.81,830800.1650.420.7014,399,2002,735,400203,300
43.02.5642.91,289300.0160.104.7216,579,1004,915,300414,500
55.02.5671.52,148300.0270.107.8616,579,1004,915,300414,500
610.02.56143.04,297300.0540.1015.7316,579,1004,915,300414,500
To calculate the drag (Cd) and inertia (Cm) coefficients, Morrison’s equation (Sumer and Fredsøe 1997) as shown later was fitted to the sampled numerical phase-averaged total force (Ftot), calculated the same way as in Eq. (12), over one time period, using a nonlinear least-square fitting technique (López and García 1991)
Ftot(t)=Fdrag(t)+Finertia(t)=12CdρlDU(t)|U(t)|+ρCmπD24U(t)t
(15)
where U(t) = phase-averaged velocity measured at the free surface, averaged at the sampling locations in Fig. 1. The least-square method minimizes the mean square error (ϵ2) between the measured and predicted forces to determine Cd and Cm and is expressed as
ϵ2=1Ni=1N(FniFpi)2
(16)
where subscripts n and p, respectively, represent the sampled (numerical) and predicted (using Morrison’s equation) values of the total force as in Eq. (15); and N = total number of discrete data points. Using ϵ2/Cd=0 and ϵ2/Cm=0, the following two equations with two unknowns Cd and Cm are obtained, which were then solved using MATLAB
fd·U4(t)+fm·U(t)·|U(t)|·U˙(t)=U(t)·|U(t)|·Fn(t)
(17)
fd·U(t)·|U(t)|·U˙(t)+fm·U˙2(t)=U˙(t)·Fn(t)
(18)
where U˙(t) = first derivative of the phase-averaged velocity at the free surface; and fd=12ρCdD and fm=π4ρCmD2. The summation limits have been omitted for ease of reading and span from 1 to N, where N is the total number of discrete intervals within a time period. N=300 was selected for all runs.
Fig. 4 shows the comparison of the variation of simulated drag and inertia coefficients with KC numbers (KC=UmaxT/D, where Umax is the maximum orbital velocity from linear wave theory at the SWL) for the array and isolated cylinder cases with the experimental values along with the β=R/KC numbers for each case. The agreement with the experimental data is good and shows that the model is able to simulate the fitted force coefficients accurately. Both Cd and Cm decrease with the KC number for the range of values tested for the β=80 case, whereas for the smaller β=30 the coefficients remain almost constant with KC showing only a reduction at KC=143. The isolated cylinder shows marginally lower force coefficients than the corresponding array cylinder, whereas there is little overall difference between the array and isolated cylinder results in the simulations, which supports the observations of Hayashi and Chaplin (2012) who concluded that the spacing between the arrays (10D) was likely large for any significant vortex interactions. This is an interesting departure from the unidirectional flow results that show at similar spacings vortex interactions are quite active and influence the shedding dynamics of the cylinders.
Fig. 4. Validation of force coefficients from wave flow experiments: comparison of drag (Cd) and inertia (Cm) predicted by the model for the central cylinder in the array and for the isolated cylinder tests with the corresponding experimental values adapted from Hayashi and Chaplin (2012) at two different β values and for a range of KC numbers
Fig. 5(a) shows the time histories (about four wave cycles) of forces acting on the cylinders located along the centerline [Fig. 5(b) for exact cylinder locations] in the direction of wave propagation for Cases 1, 2, 5, and 6. Results for Case 3 are similar in trend to Cases 1 and 2, whereas those of Case 4 are similar in trend to Cases 5 and 6 and are not shown here. Although the RMS force magnitudes show little overall variation down the flume, the higher KC number cases show a greater variation, with the cylinders in the front of the flume (3 and 23) experiencing higher forces than those toward the back (83 and 103). There is a marginal increase in the force experienced by the central cylinder (53), which may be caused by a weak reflection from the back of the channel causing a partial standing wave antinode at this point. The isolated cylinder consistently experiences higher forces than the array cylinder at the same location. The column of panels on the right shows the force histories averaged over a single phase and time lagged over a single time window. It also clearly shows the variability in the force histories as well as the effect of the nonlinearity of the wave condition, particularly at higher KC numbers in which a clear inflection point is visible below the SWL line. The crest is more peaked than the trough at large KC numbers with the force direction showing an asymmetry with the wave phase.
Fig. 5. Variation of depth-integrated forces along the flume centerline: (a) force time histories and corresponding phase-averaged forces for the cylinders marked by dark circles in (b); (b) location of all cylinders in the flume
The top row of panels in Fig. 6 shows the time series of pressure (Fp), viscous horizontal (Fv,x), viscous vertical (Fv,z), and the total force (Ftot=Fp+Fv,x) over about four wave cycles for Cases 1 and 6. The bottom row of panels shows the time series of the corresponding moments about the stem base. The surface elevation (η) is also plotted on a second y-axis on the right. The viscous forces are almost in phase with the surface elevation and, consequently, lead the pressure force by almost 90° at the lower KC number. This phase difference is, however, not so pronounced at the higher KC number. In terms of force magnitudes, it is found that at low KC numbers the viscous horizontal force contributes significantly to the total force, whereas at higher KC the contribution is negligible. The viscous vertical force is also higher than the viscous horizontal force at KC=9.1. The viscous forces appear to be positive mostly with sharp peaks and shallow troughs. The phase difference between the pressure force and the surface elevation is near 90° at KC=9.1, suggesting the contribution of the inertia effect, whereas the forces are almost in phase with the surface elevation at the higher KC. The addition of the viscous force to the pressure force decreases the phase difference somewhat between the total force and the surface elevation. The moments follow a similar trend to the forces except a change in scale as expected.
Fig. 6. Comparison of forces and moments at low and high KC numbers: moments are measured about the cylinder base; the dark dotted line signifies the end of each wave period
The upper row of panels in Fig. 7 shows the fitted drag and inertia forces according to Morrison’s equation at two representative KC numbers. The drag force and inertia force are 90° out of phase with each other because of the nature of Morrison’s equation with the inertia force leading the drag force. It is found that at lower KC=9.1, the inertia force is significant in magnitude and is even greater than the drag force, whereas at the higher KC the drag force is much larger. This causes a temporal spreading of the total force for the lower KC number and results in a phase difference with the surface elevation. At the higher KC number in which the drag force is dominating and is almost in phase with the surface elevation, the total force follows a similar trend. The lower row of panels in Fig. 7 shows the time-lagged forces. The time lagging was done by calculating the correlation between the surface elevation time series and the force’s time series and finding out the lead or lag and shifting the force’s time series accordingly. It is seen that the effect of time lagging causes the predicted total forces to be almost double that of the measured total forces for the lower KC case, whereas the total force increases only slightly for the higher KC case. In most large-scale models it is a convenient practice to only consider time- and depth-averaged drag forces based on some a priori assumed drag coefficient value and to neglect the effect of the inertia forces for calculating the total forces acting on model plants because inertia forces are conservative (for linear wave assumption only). Based on the findings here it appears that although it may be fine to neglect the inertia effect at high KC numbers because they are too small, the same cannot be said at low KC numbers in which an overestimation of total forces might result by assuming drag contribution only.
Fig. 7. Effect of time-lagging fitted force curves at low and high KC numbers: the fitted force curves according to Morrison’s equation are shown along with the additive effect on the net measured force if they are assumed to have zero time lag between them; note the difference in coordinate ranges for the force axes for the two cases
In Fig. 8, the vertical variation of forces [Fig. 8(a)] and vertical variation of corresponding moments [Fig. 8(b)] acting on the central cylinder (53) in Fig. 5(b) for four representative cases (Cases 1, 2, 5, and 6) are shown. The crest or high-water level (HWL), SWL, and trough or low-water level (LWL) are marked by black dashed lines as well as the central zero line passing through the zero mark on the X-axis. For the force calculations, equations similar to Eq. (12) were used except the depth-integrating operation. Calculation of forces around the air-water interface can be difficult in VOF models because of the absence of a well-defined interface, thus to only consider fluid-filled regions, pressure (p*), and effective viscosity (τeff) values from cells with void fractions equal to or greater than 0.5 were used in the equations. Maximum and minimum forces and corresponding moments about the base of the cylinder on the flume bed signify the maximum positive and maximum negative forces acting on the cylinder over the entire wave record, whereas the RMS and mean quantities have their usual meaning. For all cases the peaks of the maximum (positive) force and moment occur just below the HWL but above the SWL. A similar phenomenon was observed by Torum (1989) who explained it by the analogy of unidirectional flow with free surface in which the pileup and increase of water level on the upstream side and the additional drawdown on the downstream face causes the pressure difference. As will be explained later in this section, a similar gradient in the mean velocity (and hence pressure) exists between the SWL and the HWL on either side of the cylinder in the wave direction due to the net positive velocity associated with the shoreward motion of the wave crest. The rising limb of the wave causes a recirculation in the horizontal plane, which spans vertically throughout between the wave crest and trough, whereas the falling limb of the wave has a recirculation primarily only below the SWL. This is also confirmed by the evidence that the minimum force (maximum of negative forces) almost vanishes between the SWL and HWL, suggesting that the recirculation region on the rising limb of the wave is stronger than the one developed during the falling limb. The peak of the RMS and minimum (maximum negative) forces and moments values occur just below the LWL. The mean values show a slight negative bias below the LWL but are entirely positive between the LWL and HWL with a peak lying at almost the same level as the maximum force. A local peak in the RMS value at that level is also noticed. A clear distinction can be seen between the higher and lower KC number cases. The lower KC number cases show a pronounced hyperbolic profile, whereas for the higher KC numbers the profiles are flatter in nature. Also interesting is the distinct boundary layer effect on the force profiles. The force profiles show a decrease within the wave bottom boundary layer, and this decrease is seen within a distance of 5–6 cm from the bed for the lower KC numbers to 2–3 cm for the higher KC numbers.
Fig. 8. Vertical variation of forces and moments on the central cylinder in the array: the dark dotted horizontal lines signify the HWL, mean water level, and LWL for each case; the central vertical dotted line signifies location of the zero in the X-axis: (a) forces for four select cases; (b) moments for four select cases
To analyze the hydrodynamics in the water column and investigate the spatial influence of the cylinder in the wave flow in the flume, the vertical and horizontal velocities at several locations in the central plane were compared with the array and isolated cases. Figs. 9 and 10 show the profiles of horizontal (Urms) and vertical (Wrms) RMS of the wave velocities at 10 locations (five on each side of the cylinder, in which distances are made nondimensionalized by the cylinder diameter) including the four locations marked in Fig. 1(d) for the array case for Wave Cases 5 (kH/2=0.027) and 6 (kH/2=0.054). Here, for points between the crest and trough the temporal RMS operation considers velocity values from points below the water surface only. The results from the other cases show similar conclusions and are not shown here. The dotted dark black line shows the theoretical velocity profile at the center of the flume from the wave flume-only case (no cylinder) calculated using the inviscid Stokes’ second-order theory (Svendsen 2006). To calculate the Stokes’ theoretical profiles, the observed water surface elevation time series for each wave case was fitted for some unknown x and H against the Stokes’ second-order equation for surface elevation as
ηstokes(t)=H2cos(kxωt)+116kH2(3coth3khcotkh)cos2(kxωt)
(19)
and the best fit x and H were substituted in the equations for horizontal (ustokes) and vertical (wstokes) orbital velocities as
ustokes(t)=ωH2coshk(z+h)sinhkhcos(kxωt)+316c(kH)2cosh2k(z+h)sinh4khcos2(kxωt)
(20a)
wstokes(t)=ωH2sinhk(z+h)sinhkhsin(kxωt)+316c(kH)2sinh2k(z+h)sinh4khsin2(kxωt)
(20b)
Fig. 9. Variation of RMS horizontal wave orbital velocities (URMS), along the flume centerline, at various distances from the cylinder center; the Stokes’ second-order profile uses the free surface elevation at the flume center (no-cylinder case): (a) Case 5; (b) Case 6
Fig. 10. Variation of RMS vertical wave orbital velocities (WRMS), along the flume centerline, at various distances from the cylinder center; the Stokes’ second-order profile uses the free surface elevation at the flume center (no-cylinder case): (a) Case 5; (b) Case 6
It was found that the horizontal and vertical velocity profiles at far-field locations |X|=10D and |X|=8.67D for the isolated cylinder case almost match the numerical results for the no-cylinder case, meaning that there is very little variation of the global wave flow field at these distances on either side of the cylinder. The array cylinder results (shown by green line) show that the velocities are slightly less than the isolated and no-cylinder cases because of the small damping effect of the cylinders in the front half of the canopy. Within |X|2D significant departure of the wave orbital horizontal RMS velocities is seen (Fig. 9). Below the wave trough or LWL the recirculation and turbulence action on either side of the cylinder decreases the wave excursions significantly compared with the no-cylinder case. The higher wave height case (Case 6) shows larger deviation caused by larger turbulence loss and thus greater wave orbital energy damping by the stem. Velocity profiles for both array and isolated cylinders in the region between the wave crest and the wave trough differ significantly on either side of the cylinder within the recirculation zone (|X|2D). This apparent anomaly is explained further in Fig. 11, which shows the mean horizontal velocities at the same locations as in Fig. 9 for the two cases, with the temporal mean velocity calculated between the trough and the crest considering points below the water surface only. Below the wave trough the mean flow is almost zero for the far-field locations (|X|=10D and |X|=8.67D), whereas within X|2D, below the wave trough, the profiles show a net positive current, and behind the cylinder a net negative current caused by the recirculation region exists on either side of the cylinder. Between the wave trough and the crest, the far-field profiles show a net positive current caused by the forward motion of the wave crest that is entirely positive, which is expected for progressive waves. Near the cylinder, within the recirculation zone, is observed a decrease of this positive current behind the cylinder in the wave direction caused by the effect of the surface vortex (and hence negative velocities) created during the forward thrust of the crest behind the cylinder in the wave direction. This also explains the existence of a free surface velocity gradient on either side of the cylinder between the trough and the crest as was mentioned earlier. The RMS horizontal velocity profiles (Fig. 9) likewise show a reduction behind the cylinder between the trough and the crest compared with the points in front of the cylinder in the wave direction. The reduction in the free surface horizontal wave velocity seems to increase with wave nonlinearity (Case 5 versus Case 6), because the greater the nonlinearity the greater is the Stokes’ drift associated with the wave orbital excursion within the trough and the crest. In other words, because the time-averaged horizontal flow between the trough and the crest behaves as a uniform flow, with a vertical variation, it creates a mean damping effect similar to that observed in uniform flow around a cylinder, whereas the flow below the trough, having predominantly wave orbital signatures, is damped similar to wave damping in submerged vegetation. This separation of the hydrodynamic regime for an emergent vegetation problem has a potentially important influence on simplified quadratic law–based formulations commonly used in large-scale models and can be a topic of future research.
Fig. 11. Vertical variation of mean horizontal velocities (U¯) along the flume centerline, at various distances from the cylinder center; the array case results are almost similar to the corresponding isolated cases and are omitted here: (a) Case 5; (b) Case 6
Vertical wave velocity RMS profiles (Fig. 10) show similar trends at far-field locations (X=10D and X=8.67D) for array, isolated, and no-cylinder cases. Within the recirculation zone (X2D) the profiles are similar, except near the free surface. In front of the cylinder, the obstruction by the cylinder causes run up on the seaward face during the forward motion of the wave crest followed by a localized recirculation vortex created below the SWL during the flow-reversal phase and impinging into the SWL. This causes a local increase in vertical velocity excursions. Behind the cylinder, because there is no impinging below the SWL of the surface vortex, the recirculation exists only when the crest is in shoreward motion causing lesser vertical velocity excursion amplitudes compared with that on the front face.
The previously mentioned free surface phenomenon is visualized in Fig. 12(a), in which the left panel represents Wave Case 5 and the right panel represents Wave Case 6. The figure shows from top to bottom, respectively, the mean horizontal, mean vertical, RMS horizontal, and RMS vertical velocities over 20 wave periods presented with the color bar on the right. The crest, mean, and trough locations are also indicated in the same figure. As explained in the line diagrams (Figs. 911), a net positive horizontal velocity flux above the SWL (top panels) is observed, which increases with increasing nonlinearity. Fig. 12(b) (bottom two rows) shows conceptually the wave height–integrated horizontal recirculation zones around the cylinder below and above the SWL during the rising limb (shoreward motion of wave crest) and falling limb (seaward motion of wave crest) of the free surface (locations shown relative to the cylinder in the top row). Thus, over a complete cycle, the excess flux above the SWL causes a uniform-flow type phenomenon with a pronounced recirculation zone behind the cylinder in the direction of wave propagation, particularly between the crest and the SWL, which makes the depth-integrated recirculation zone between the trough and the crest stronger in the shoreward direction of the cylinder. The shear effect of the flux also extends downward below the SWL. The mean vertical velocities in Fig. 12(a) show an asymmetrical signature with a net negative zone between the crest and the trough behind the cylinder and one between the trough and the SWL in front of it. This is caused by the impinging action of the surface vortex into the SWL during the flow reversal in the wave crest. The RMS horizontal velocity plots indicate the bulk of the free surface energy is damped between the SWL and wave crest, whereas below the trough the damping occurs on both sides with similar magnitudes similar to wave damping around submerged bluff bodies.
Fig. 12. Hydrodynamics between the crest and the trough: (a) vertical distribution of mean (U¯ and W¯) and root mean square (URMS and WRMS) velocities between the trough (ηmin) and the crest (ηmax) for Case 5 and Case 6 near the emergent cylinder (isolated); (b) conceptual sketch of horizontal recirculation zones below and above SWL at different wave phases (LR is the length of the recirculation zone)
Vortex growth, shedding, and subsequent force variations near the free surface from an emergent surface-piercing cylinder has been a topic of active research (Kjeldsen et al. 2011; Torum 1989; Quetey et al. 2004; Barlas 2012). However, detailed visualization of the vortex shedding near the surface for an emergent cylinder has not been previously reported in the literature. Fig. 13 presents the 3D vortex visualizations from the isolated cylinder run corresponding to Case 5, as in Table 2, as a representative of the free surface damping phenomenon explained previously. Vortex cores are represented by the Q-criterion (Hunt et al. 1988), and isosurfaces are drawn at constant Q=10 with coloring done by vorticity (red-green-blue scale) about the z-axis with red representing a clockwise rotating vortex (positive) when viewed in the z-direction from the origin, and blue represents a counterclockwise (negative) vortex. The Q-criterion is mathematically represented as
Q=12(Ω2S2)
(21)
where Ω=tr[ΩΩt]1/2; and S=tr[SS2]1/2 = asymmetric and symmetric parts of u¯ defined as
S=12(u¯+(u¯)t)
(22a)
Ω=12(u¯(u¯)t)
(22b)
Fig. 13. Vortex shedding from an isolated cylinder (Wave Case 5) at various stages of a wave period: vortex isosurfaces are drawn according to the Q-criterion and colored by vorticity (ω) about the z-axis: (a) t = T/8; (b) t = T/4; (c) t = 3T/8; (d) t = T/2; (e) t = 5T/8; (f) t = 3T/2; (g) t = 7T/8; (h) t = T
A vortex region is said to exist in a flow field if Q>0 in the cells comprising that region and may be thought of as local regions having a greater rotation rate than strain rate. Figs. 13(a–h) show the instantaneous vortex isosurfaces generated near the free surface from the cylinder at eight equal intervals of time along a representative wave period (T). The figures have been magnified to show the region near the vicinity of the free surface with black lines marking the crest elevation or HWL, the mean water level or SWL, and trough elevation or LWL. At t=T/8 [Fig. 13(a)], the free surface is accelerating and descending downward, as shown by the arrow direction. Surface Vortex A, having its axis almost parallel to the SWL and forming a circular arc around the cylinder, is the remnant of the vortex shed in the preceding cycle, whereas two counterrotating vortex pairs (B-B’) that also shed in the preceding cycle are seen. A violent free surface also exists near the cylinder-water interface as well as strong isolated vortex streaks formed just below the free surface. At t=T/4 [Fig. 13(b)], as the downward acceleration of the free surface has reached a peak, Vortex A has formed almost a helical shape, is impinging below the free surface, and appears to be dissipating. This vortex aligns itself mostly between the SWL and somewhat below the LWL and is responsible for the increase in vertical velocity excursions in this zone. The counterrotating pair has weakened and is ready to be flipped over the cylinder. The rotation within the boundary layer also has changed direction with a weak vorticity structure developing. At t=3T/8 [Fig. 13(c)], as the downward motion of the free surface decelerates Vortex A is seen to be dissipating further and moving downward in a helical pattern. B-B’ vortices have flipped sides at this point and also have been pushed down with the top part almost dissipating under the strong vertical velocities. At t=T/2 [Fig. 13(d)], as the wave is about to change direction, Vortex A has almost dissipated with only weak helical signatures surrounding the cylinder. B-B’ has completely dissipated from the window at this stage. At t=5T/8 [Fig. 13(e)], as the free surface accelerates upward, the counterrotating vortex pair (C-C’) that was developing within the boundary layer since t=T/4 has now shed completely. It is worth noticing that the vortex is almost gone except for a few streaks near the free surface. The rise of the free surface actually causes a stretching of the counterrotating pair and they reach very near to the free surface. No prominent surface vortices are visible, unlike what was seen at t=T/8. At t=3T/2 [Fig. 13(f)], the free surface has reached the highest peak of upward acceleration and causes significant stretching and thinning of the C-C’ pair. Surface Vortex A has started shedding and is present only as a few streaks. The recirculation of this zone associated with the rising limb gives rise to the negative velocity zone behind the cylinder in the wave direction and will cause dampening of the surface energy in the subsequent two-phase instants. C-C’ is about to flip over the cylinder, and new counterrotating vortex signatures can be seen developing within the boundary layer, which will eventually become the B-B’ pair in the subsequent cycle. At t=7T/8 [Fig. 13(g)], the free surface rise has decelerated the detachment of the C-C’ vortex pair, and formation of a new A vortex is seen. Stronger vortices are formed under the free surface on the lee side of the cylinder during the rising limb of the wave than during the falling limb. This causes a stronger recirculation region in the wave direction and, consequently, higher positive forces on the cylinder under the SWL during the rising cycle of the wave than during the falling limb. It also explains the asymmetrical surface velocities observed previously on either side of the cylinder. Finally, at t=T [Fig. 13(h)] C-C’ is completely dissipated while Vortex A aligns itself in a circular arc around the cylinder near the free surface.

Conclusions

In this paper the unidirectional flow over submerged vegetation and free surface wave flow over emergent vegetation were simulated using the dynamic Smagorinsky LES turbulence closure scheme to validate and demonstrate the capabilities of the model when applied to vegetated flows. The vegetation was idealized as an array of rigid cylinders. The main conclusions from this study can be summarized as follows:
1.
Low-resolution LES, with adaptive mesh refining around the cylindrical stems, is able to produce fairly accurate flow descriptions of the first- and second-order velocity quantities, comparable to results of other high-resolution studies. The present simulation results matched the mean and turbulent velocities within and above the canopy from experiments of Liu et al. (2008) with reasonable accuracy. Further, bulk drag coefficient calculated from the depth-integrated, direct force measurements around the cylinders match the value predicted by the empirical equations of Ghisalberti and Nepf (2004) and Nepf (1999), indicating that optimal-resolution LES simulations yielding acceptable velocity distribution within the canopy can be used to evaluate drag coefficients within a margin of error. In terms of computational effort, the present simulations run approximately 25 times faster than the high-resolution LES experiments of Stoesser et al. (2009) in comparable hardware. It was found that a minimum of 60 grid points was required on the cylinder circumference to accurately simulate the drag and inertia forces and yet maintain optimal resolution away from the cylinder in the free zones within the canopy in which the grid was coarsened, whereas the vertical aspect ratio (ΔxΔz) can be increased up to 2.5 without compromising the model results.
2.
Force coefficients calculated by fitting Morrison’s equation to the observed forces in the wave flow experiments matched well with the experiments of Hayashi and Chaplin (2012). No significant differences in force measurements were found between the isolated and array cylinder, which is an indication that the hydrodynamics around the central array cylinder is similar to that of an isolated cylinder for the spacing chosen.
3.
Vertical variation of mean and RMS wave velocity profiles at several locations along the central vertical plane on either side of the cylinder at the array center were compared with those at the same locations around the isolated cylinder and that in the wave flume without a cylinder. Notable differences were seen in the velocities on either side of the cylinder in the zone between the wave trough and wave crest, and this difference seems to increase with wave nonlinearity. The shoreward motion of the wave energy (during the rising limb of the water surface) associated with the wave crest impacting the cylinder creates a localized recirculation zone spanning throughout between the wave trough and the wave crest on the shoreward side of the cylinder, whereas the seaward motion of the wave crest (during the falling limb of the water surface) creates a recirculation zone located only between the wave trough and the SWL on the seaward side of the cylinder. Also, the seaward recirculation zone between the trough and the SWL is weaker compared with the shoreward one (during the shoreward motion of the crest) for waves with greater nonlinearity caused by Stokes’ drift. Thus, over one representative period a net reduction of mean and RMS velocities is seen shoreward of the cylinder in the vicinity of the SWL with respect to an equidistant point in front of the cylinder. For the cases investigated, this effect was most prominent within two diameters from the cylinder center. This also verifies that the array density was too sparse for notable hydrodynamic interaction effects to be observed for the array cylinders. It also explains the agreement of the drag coefficient values with the isolated cylinder case.
4.
Vortex shedding was found to be largely localized around a one- to two-diameter distance on either side of the cylinder, and no appreciable vortex interaction with the neighboring cylinders was noted. The vortex evolution pattern was found to be different between the zone above the wave trough from the rest of the cylinder. This is because the net horizontal wave orbital velocities, between the SWL and the crest, always point in the direction of wave propagation during the forward motion resulting in a stronger recirculation shoreward of the cylinder between the trough and the crest over a wave period. The vertical orbital velocity amplitudes show little deviation from Stokes’ velocities, except near the SWL, in which the run-up effect on the seaward face of the cylinder causes a local increase of the vertical velocity amplitude. This also indicates that the horizontal velocity components are damped more than the vertical component for the majority of the cylinder height.
5.
Vertical profile of maximum force and moment on the cylinder suggests the peak force occurs somewhere between the SWL and the HWL and agrees with the experimental evidence in the literature.

Notation

The following symbols are used in this paper:
B
width of domain (m);
D
cylinder diameter (m);
g
acceleration due to gravity (9.81m/s2);
H
mean wave height (m);
h
water depth or submerged height of the domain (m);
k=2π/L
wave number (1/m);
L=L0tanh(kh)
wavelength of wave solved iteratively using the dispersion relationship (m);
L0=gT2/2π
deep-water wavelength (m);
Rcanopy=Uc¯Dν
canopy Reynolds number;
Rchannel=UbBν
channel Reynolds number;
RD=UbDν
cylinder Reynolds number;
T
wave period (s);
t
time (s);
Ub
bulk velocity in the channel (m/s);
U*
shear velocity (m/s);
Uc¯
averaged velocity in the canopy layer (m/s);
x, y, z
generic x, y, and z coordinates;
y+=U*y/nueff
nondimensional wall units, y is measured along local normal to the wall;
Δx, Δy, Δz
cell sizes in X-, Y-, and Z-directions (m);
μa=ρaνa
dynamic molecular viscosity of water (103kg/m3);
μw=ρwνw
dynamic molecular viscosity of water;
νa
molecular kinematic viscosity of air (0.148×106m2/s);
νw
molecular kinematic viscosity of water (1×106m2/s);
ρa
density of air (1kg/m3);
ρw
density of water (103kg/m3); and
ω=2π/T
wave angular velocity (rad/s).

Acknowledgments

The study was supported in part by the U.S. National Science Foundation (DMS-1115527, DMS-1115546 and CCF-1539567) and the Louisiana Coastal Protection and Restoration Authority (CPRA) Applied Research Program. High-performance computational resources for this research have been provided by the Center for Computation and Technology at Louisiana State University.

References

Asano, T., Tsutsui, S., and Sakai, T. (1988). “Wave damping characteristics due to seaweed.” Proc., 35th Coast. Eng. Conf. in Japan, JSCE, Japan, 138–142 (in Japanese).
Augustin, L., Irish, J., and Lynnet, P. (2009). “Laboratory and numerical studies of wave damping by emergent and near-emergent wetland vegetation.” Coast. Eng., 56(3), 332–340.
Barlas, B. (2012). “Interactions of waves with an array of tandem placed bottom-mounted cylinders.” J. Mar. Sci. Technol., 20(1), 103–110.
Berberović, E., van Hinsberg, N. P., Jakirlić, S., Roisman, I. V., and Tropea, C. (2009). “Drop impact onto a liquid layer of finite thickness: Dynamics of the cavity evolution.” Phys. Rev. E, 79(3), 036306-1–036306-15.
Chakrabarti, A., and Smith, H. (2011). “Investigation of turbulent structure in emergent vegetation under wave forcing.” Proc., Coastal Sediments, ASCE, Reston, VA.
Chen, Q., and Zhao, H. (2012). “Theoretical models for wave energy dissipation caused by vegetation.” J. Eng. Mech., 221–229.
Cui, J., and Neary, V. (2002). “Large eddy simulation (LES) of fully developed flow through vegetation.” Proc., 5th Int. Conf. Hydroinformatics, International Association of Hydraulic Research, Cardiff, U.K.
Dalrymple, R., Kirby, J., and Hwang, P. (1984). “Wave diffraction due to areas of energy dissipation.” J. Waterway, Port, Coastal, Ocean Eng., 67–79.
Dunn, C. J., López, F., and García, M. H. (1996). “Mean flow and turbulence in a laboratory channel with simulated vegetation.”, U.S. Army Corps of Engineers, Washington, DC.
Engsig-Karup, A. P. (2006). “Unstructured nodel DG-FEM solution of high-order Boussinesq-type equations.” Ph.D. thesis, Technical Univ. of Denmark, Kongens Lyngby, Denmark.
Fairbanks, J., and Diplas, P. (1998). “Turbulence characteristics of flows through partially and fully submerged vegetation.” Proc., Wetlands Eng. River Restoration Conf., ASCE, Reston, VA.
Ferziger, J., and Peric, M. (2001). Computational methods for fluid dynamics, 3rd Ed., Springer, Paris.
Fischer, A. T., Stoesser, T., Bates, P., and Olsen, N. R. B. (2001). “3D numerical modelling of open-channel flow with submerged vegetation.” IAHR J. Hydraul. Res., 39(3), 303–310.
Germano, M., Piomelli, U., Moin, P., and Cabot, W. H. (1991). “A dynamic subgrid-scale eddy viscosity model.” Phys. Fluids, 3(7), 1760.
Ghisalberti, M., and Nepf, H. M. (2002). “Mixing layers and coherent structures in vegetated aquatic flows.” J. Geophys. Res., 107(C2), 3-1–3-11.
Ghisalberti, M., and Nepf, H. M. (2004). “The limited growth of vegetated shear layers.” Water Resour. Res., 40, W07502.
Ghisalberti, M., and Nepf, H. M. (2006). “The structure of the shear layer in flows over rigid and flexible canopies.” Environ. Fluid Mech., 6(3), 277–301.
Hayashi, K., and Chaplin, J. (2012). “Wave damping by an array of circular cylinders and group of model plants.” Proc., 9th Int. Symp. on Ecohydraulics, Univ. of Natural Resources and Life Sciences, IAHR, Vienna, Austria.
Hirt, C., and Nichols, B. (1981). “Volume of fluid (VOF) method for dynamics of free boundaries.” J. Comp. Phys., 39(1), 201–225.
Hu, Z., Suzuki, T., Zitman, T., Uittewaal, W., and Stive, M. (2014). “Laboratory study on wave dissipation by vegetation in combined currentwave flow.” Coast. Eng., 88, 131–142.
Hunt, J. C. R., Wray, A. A., and Moin, P. (1988). “Eddies, stream, and convergence zones in turbulent flows.”, Center for Turbulence Research, Stanford, CA.
Isaacson, M., Subbiah, K., and Baldwin, J. (1991). “Force coefficient estimation from random wave data.” Proc., 1st Int. Offshore and Polar Engineering Conf., International Society of Offshore and Polar Engineers, Mountain View, CA.
Jacobsen, N. G., Fuhrman, D. R., and Fredsøe, J. (2012). “A wave generation toolbox for the open-source cfd library: Openfoam®.” Int. J. Numer. Meth. Fluids, 70(9), 1073–1088.
Jadhav, R. S., and Chen, Q. (2013). “Probability distribution of wave heights attenuated by salt marsh vegetation during tropical cyclone.” Coast. Eng., 82, 47–55.
Jadhav, R. S., Chen, Q., and Smith, J. M. (2013). “Spectral distribution of wave energy dissipation by salt marsh vegetation.” Coast. Eng., 77, 99–107.
Kjeldsen, S., Torum, A., and Dean, R. (2011). “Wave forces on vertical piles caused by 2- and 3-dimensional breaking waves.” Proc. Coast. Eng., 1(20), 1929–1942.
Kobayashi, N., Baichle, A. W., and Asano, T. (1993). “Wave attenuation by vegetation.” J. Waterway. Port, Coastal, Ocean Eng., 30–48.
Kortlever, W. R. (1994). “Wave attenuation by using reed for bank protection.”, Delft Univ. of Tech, Delft, Netherlands.
Kouwen, N., Unny, T., and Hill, H. (1969). “Flow retardance in vegetated channels.” J. Irrig. Drain. Div., 95(2), 329–342.
Liu, D., Chen, Q., and Wang, Y. (2011). “Spectral element modeling of sediment transport in shear flows.” Comput. Meth. Appl. Mech. Eng., 200(17–20), 1691–1707.
Liu, D., Diplas, P., Fairbanks, J. D., and Hodges, C. C. (2008). “An experimental study of flow through rigid vegetation.” J. Geophys. Res., 113, 1–16.
Liu, D., Zheng, Y.-L., and Chen, Q. (2015). “Grain-resolved simulation of micro-particle dynamics in shear and oscillatory flows.” Comput. Fluids, 129–141.
López, F., and García, M. H. (1997). “Open channel flow through simulated vegetation: Turbulence modeling and sediment transport.”, U.S. Army Corps of Engineers, Waterways Experiment Station, and Dept. of Civil Engineering, Washington, DC.
López, F., and García, M. H. (2001). “Mean flow and turbulence structure of open-channel flow through non-emergent vegetation.” J. Hydraul. Eng., 392–402.
Lowe, R., Koseff, J. R., Monismith, S., and Falter, J. (2005a). “Oscillatory flow through submerged canopies: Part 1: Velocity structure.” J. Geophys. Res., 110(C10016), 1–17.
Lowe, R., Koseff, J. R., Monismith, S., and Falter, J. (2005b). “Oscillatory flow through submerged canopies: Part 2: Canopy mass transfer.” J. Geophys. Res., 110(C10017), 1–14.
Ma, G., Kirby, J. T., Su, S.-F., Figlus, J., and Shi, F. (2013). “Numerical study of turbulence and wave damping induced by vegetation canopies.” Coast. Eng., 80, 68–78.
Marsooli, R., and Wu, W. (2014). “Numerical investigation of wave attenuation by vegetation using a 3D {RANS} model.” Adv. Water Resour., 74, 245–257.
MATLAB version 8.2.0.29 [Computer software]. MathWorks, Natick, MA.
Maza, M., Lara, J., and Losada, I. (2013). “A coupled model of submerged vegetation under oscillatory flow using Navier–Stokes equations.” Coast. Eng., 80, 16–34.
Mendez, F. J., and Losada, I. J. (1999). “Hydrodynamics of a vegetation field induced by wind waves.” J. Geophys. Res., 104(C8), 18383–18396.
Mendez, F. J., and Losada, I. J. (2002). “An empirical model to estimate the propagation of random breaking and nonbreaking waves over vegetation fields.” Coast. Eng., 51(2), 103–118.
Naot, D., Nezu, I., and Nakagawa, H. (1996). “Hydrodynamic behavior of partially vegetated open channels.” J. Hydraul. Eng., 625–633.
Neary, V. S. (2003). “Numerical solution of fully developed flow with vegetative resistance.” J. Eng. Mech., 558–563.
Nepf, H. M. (1999). “Drag, turbulence, and diffusion in flow through emergent vegetation.” Water. Resour. Res., 35(2), 479–489.
Nepf, H. M., and Ghisalberti, M. (2009). “Flow and transport in channels with submerged vegetation.” Acta Geophys., 56(3), 753–777.
Palau-Salavador, G., Stoesser, T., Froehlich, J., Kappler, M., and Rodi, W. (2010). “Large eddy simulations and experiments of flow around finite-height cylinders.” Flow Turbul. Combust., 84(2), 239–275.
Quetey, P., Visonneau, M., and Ferrant, P. (2004). “Numerical investigation of wave interaction with a fixed vertical circular cylinder.”, OnePetro, Richardson, TX.
Sagaut, P. (1998). Large eddy simulation of incompressible flows, 3rd Ed., Springer, Berlin.
Seng, S., Jensen, J., and Pedersen, P. (2012). “Numerical prediction of slamming loads.” J. Eng. Marit. Environ., 226, 120–134.
Shimizu, Y., and Tsujimoto, T. (1994). “Numerical analysis of turbulent open-channel flow over vegetation layer using a k-turbulence model.” J. Hydrosci. Hydraul. Eng., 11(2), 57–67.
Spalding, D. (1961). “A single formula for the law of the wall.” J. Appl. Mech., 28(3), 455–458.
Stoesser, T., Palau-Salvador, G., Rodi, W., and Diplas, P. (2009). “Large eddy simulation of turbulent flow through submerged vegetation.” Transp. Porous Media, 78(3), 347–365.
Sumer, B. M., and Fredsøe, J. (1997). Hydrodyn. Cylindrical Struct., World Scientific, Singapore.
Svendsen, I. A. (2006). Introduction to nearshore hydrodynamics, 1st Ed., World Scientific: Advanced Series on Ocean Engineering, Singapore.
Tanino, Y., Nepf, H., and Kulis, P. S. (2005). “Gravity currents in aquatic canopies.” Water. Resour. Res., 41(12).
Torum, A. (1989). “Wave forces on pile in surface zone.” J. Waterway, Port, Coastal, Ocean Eng., 547–565.
Tsujimoto, T., Shimizu, T., and Okada, T. (1991). “Turbulent structure of flow over rigid vegetation-covered bed in open channels.”, Hydraulic Laboratory, Kanazawa Univ., Japan.
Weller, H., and Tabor, G. (1998). “A tensorial approach to computational continuum mechanics object-oriented techniques.” Comput. Phys., 12(6), 620–631.
Wu, W., Zhang, M., Ozeren, Y., and Wren, D. (2013). “Analysis of vegetation effect on waves using a vertical 2D {RANS} model.” J. Coast. Res., 287(2), 383–397.
Zhao, H., and Chen, Q. (2014). “Modeling attenuation of storm surge over deformable vegetation: Methodology and verification.” J. Eng. Mech., 04014090.
Zhu, L., and Chen, Q. (2015). “Numerical modeling of surface waves over submerged flexible vegetation.” J. Eng. Mech., A4015001.

Information & Authors

Information

Published In

Go to Journal of Engineering Mechanics
Journal of Engineering Mechanics
Volume 142Issue 8August 2016

History

Received: Dec 15, 2014
Accepted: Jan 14, 2016
Published online: Apr 1, 2016
Published in print: Aug 1, 2016
Discussion open until: Sep 1, 2016

Authors

Affiliations

Agnimitro Chakrabarti [email protected]
Graduate Research Assistant, Dept. of Civil and Environmental Engineering, Louisiana State Univ., 3418 Patrick F. Taylor Hall, Baton Rouge, LA 70803. E-mail: [email protected]
Professor, Dept. of Civil and Environmental Engineering, Louisiana State Univ., 3418 Patrick F. Taylor Hall, Baton Rouge, LA 70803 (corresponding author). E-mail: [email protected]
Heather D. Smith [email protected]
Lecturer, Dept. of Civil Engineering, Univ. of New Hampshire, W139 Kingsbury Hall, 33 Academic Way, Durham, NH 03824. E-mail: [email protected]
Associate Professor, Dept. of Mathematics, Statistics and Mechanical Engineering, Louisiana Tech Univ., Bogard Hall 226, Ruston, LA 71272. E-mail: [email protected]

Metrics & Citations

Metrics

Citations

Download citation

If you have the appropriate software installed, you can download article citation data to the citation manager of your choice. Simply select your manager software from the list below and click Download.

Cited by

View Options

Media

Figures

Other

Tables

Share

Share

Copy the content Link

Share with email

Email a colleague

Share