Fluvial bedforms generate a turbulent wake that can impact suspended-sediment settling in the passing flow. This impact has implications for local suspended-sediment transport, bedform stability, and channel evolution; however, it is typically not well-considered in geomorphologic models. Our study uses a three-dimensional OpenFOAM hydrodynamic and particle-tracking model to investigate how turbulence generated from bedforms and the channel bed influences medium sand-sized particle settling, in terms of the distribution of suspended particles within the flow field and particle-settling velocities. The model resolved the effect of an engineered bedform, which altered the flow field in a manner similar to a natural dune. The modelling scenarios alternated bed morphology and the simulation of turbulence, using detached eddy simulation (DES), to differentiate the influence of bedform-generated turbulence relative to that of turbulence generated from the channel bed. The bedform generated a turbulent wake that was composed of eddies with significant anisotropic properties. The eddies and, to a lesser degree, turbulence arising from velocity shear at the bed substantially reduced settling velocities relative to the settling velocities predicted in the absence of turbulence. The eddies tended to advect sediment particles in their primary direction, diffuse particles throughout the flow column, and reduced settling likely due to production of a positively skewed vertical-velocity fluctuation distribution. Study results suggest that the bedform wake has a significant impact on particle-settling behaviour (up to a 50% reduction in settling velocity) at a scale capable of modulating local suspended transport rates and bedform dynamics.
Journal of Geophysical Research: Earth Surface, 2020
Current understanding of the role that dunes play in controlling bar and channel‐scale processes and river morphodynamics is incomplete. We present results from a combined numerical modeling and field monitoring study that isolates the impact of dunes on depth‐averaged and near‐bed flow structure, with implications for morphodynamic modeling. Numerical simulations were conducted using the three‐dimensional computational fluid dynamics code OpenFOAM to quantify the time‐averaged flow structure within a 400 m × 100 m channel using digital elevation models (DEMs) for which (i) dunes and bars were present within the model and (ii) only bar‐scale topographic features were resolved (dunes were removed). Comparison of these two simulations shows that dunes enhance lateral flows and reduce velocities over bar tops by as much as 30%. Dunes influence the direction of modeled sediment transport at spatial scales larger than individual bedforms due to their effect on topographic steering of the...
1] Fluvial channels present bed forms such as dunes and ripples that alter instantaneous hydrodynamics parameters such as flow velocities, water surface profiles, bed shear stresses, and Reynolds stresses and create turbulent coherent structures that are significantly different from those presented in flat bed conditions. It is known that LES-based models are more suitable than RANS models to reproduce the complex hydrodynamics around bed forms. Herein, a LES model is applied to describe the mean and turbulent flow structure under superimposed bed forms. Three cases were simulated: RUN I (train of ripples), RUN II (superimposed bed forms), and RUN III (amalgamated bed forms). The LES modeling was performed using a free surface condition to allow the model to develop undulations and boils on the water surface caused by effect of the bed forms. Some important conclusions from this study are: the division of high and low shear stresses on the stoss side of the dune, the progression of the flow field topology from RUN I and RUN III, and the type of turbulent coherent structures found in each stage. The region of high shear stresses was related to turbulence production, in which the streamwise velocity fluctuations (where strips structures are related to streaks) were associated to the modification of the bed morphology. The turbulence Horseshoes Vortices (THV) were more frequent in RUN I than in the other two cases (where streamwise rolls were more frequent). Finally, the frequency of the bursting events increased from RUN I to RUN II and decreased from RUN II to RUN III. Implications of detailed hydrodynamics into bed forms processes are also presented and discussed.
Understanding sediment transport process requires adequate knowledge of the mechanism of grains motion which is primarily controlled by flow characteristics including the distribution of time-averaged streamwise velocities, Reynold shear stress distributions as well as the turbulence of flow. Knowledge of the velocity profile in both clear and sediment-laden flows provide clues to understanding this sediment transport mechanism. This paper presents the characterisation of turbulent flow velocity profile based on the physical expression of the mixing length theory as originally proposed by Prandtl, O'Brien and Bagnold, for the prediction of flow interaction with suspended sediment grains. The study utilises the most current flow velocity sampling technology to directly sample flow velocity fluctuations in six cases of open channel flume experiments to characterise the turbulent velocity profile and ascertain the turbulence model's relevance and continuous application in solving sediment grain transport problems. With over 30,000 flow velocity data generated, the analysis demonstrates that, in all six clear water turbulent flows cases investigated, time-averaged velocity versus height is defined in the vicinity of the flow bed by the logarithmic law and well approximated by the turbulence model. Also, modelled and measured vertical streamwise velocities show a significant positive relationship with an R-squared value of almost unity.
Flows with high suspended sediment concentrations are common in many sedimentary environments, and their flow properties may show a transitional behaviour between fully turbulent and quasi-laminar plug flows. The characteristics of these transitional flows are a function of both clay concentration and type as well as the applied fluid stress. This paper investigates the behaviour of rapidly decelerated to steady flows that contain a mixture of sand, silt and clay, and explores the effect of different clay (kaolin) concentrations on the dynamics of flow over a mobile bed, and the bedforms and stratification produced. Experiments were conducted in a recirculating slurry flume capable of transporting high clay concentrations. Ultrasonic Doppler velocity profiling was used to measure the flow velocity within these concentrated suspension flows. The development of current ripples under decelerated flows of differing kaolin concentration was documented and evolution of their height, wavelength and migration rate quantified. This work confirms past work over smooth, fixed beds, which showed that, as clay concentration rises, a distinct sequence of flow types is generated: turbulent flow (TF), turbulenceenhanced transitional flow (TETF), lower transitional plug flow (LTPF), upper transitional plug flow (UTPF) and a quasi-laminar plug flow (QLPF). Each of these flow types produces an initial flat bed upon rapid flow deceleration, followed by reworking of these deposits through the development of current ripples during the subsequent steady flow in TF, TETF and LTPF. The initial flat beds are structureless, but have diagnostic textural properties, caused by differential settling of sand, silt and cohesive mud, which forms characteristic bipartite beds that initially consist of sand overlain by silt or clay. As clay concentration in the formative flow increases, ripples first increase in mean height and wavelength under TETF and LTPF regimes, which is attributed to the additional turbulence generated under these flows that subsequently causes greater leeside erosion. As clay concentration increases further from a LTPF, ripples cease to exist under the UTPF and QLPF conditions investigated herein. This disappearance of ripples appears due to both turbulence suppression at higher clay concentrations, as well as the increasing shear strength of the bed sediment that becomes more difficult to erode as clay concentration increases. The stratification within the ripples formed after rapid deceleration of the transitional flows reflects the availability of sediment from the bipartite bed. The exact nature of the ripple cross-stratification in these flows is a direct function of the duration of the formative flow and the texture of the initial flat bed, and ripples do not to form in cohesive flows with a Reynolds number smaller than ~12,000. Examples are given of how the unique properties of the current ripples and plane beds, developing below decelerated transitional flows, could aid in the interpretation of depositional processes in modern and ancient sediments. This includes a new model for hybrid beds that explains their formation in terms of a combination of vertical grain-size segregation and longitudinal flow transformation.
1] A series of flume experiments were conducted in a large experimental channel at the St. Anthony Falls Laboratory to understand the coupled dynamics of flow and bed forms above the sediment-water interface. Simultaneous high resolution measurements of velocity fluctuations, bed elevations and sediment flux at the downstream end of the channel, were made for a range of discharges. The probability density functions (pdfs) of bed elevation increments and instantaneous Reynolds stress reveal a power law tail behavior and a wavelet cross-correlation analysis depicts a strong dependence of these series across a range of scales, indicating a feedback between bed form dynamics and near-bed turbulence. These results complement our previous findings in which the signature of bed form evolution on the near-bed velocity fluctuations was confirmed via the presence of a spectral gap and two distinct power law scaling regimes in the spectral density of velocity fluctuations. We report herein a strong asymmetry in the probability distribution of bed elevation increments and instantaneous Reynolds stresses, the latter being further analyzed and interpreted via a quadrant analysis of velocity fluctuations in the longitudinal and vertical directions. We also report the presence of intermittency (multifractality) in bed elevation increments and interpret it, in view of the asymmetric nature of the pdfs, as the result of scale coupling. In other words, the geometric asymmetry at the bed form scale gets transferred down to a probabilistic asymmetry at all smaller scales indicating a local anisotropy in the energy transfer. Finally, we propose a predictive relationship between bed form averaged sediment transport rates and bed form averaged instantaneous Reynolds stress and validate it using our experimental data.
Detailed measurements of flow velocity and its turbulent fluctuation were obtained over fixed, two-dimensional dunes in a laboratory channel. Laser Doppler anemometry was used to measure the downstream and vertical components of velocity at more than 1800 points over one dune wavelength. The density of the sampling grid allowed construction of a unique set of contour maps for all mean flow and turbulence parameters, which are assessed using higher moment measures and quadrant analysis. These flow field maps illustrate that: (1) the time-averaged downstream and vertical velocities agree well with previous studies of quasi-equilibrium flow over fixed and mobile bedforms and show a remarkable symmetry from crest to crest; (2) the maximum root-mean-square (RMS) of the downstream velocity values occur at and just downstream of flow reattachment and within the flow separation cell; (3) the maximum vertical RMS values occur within and above the zone of flow separation along the shear layer and this zone advects and diffuses downstream, extending almost to the next crest; (4) positive downstream skewness values occur within the separation cell, whereas positive vertical skewness values are restricted to the shear layer; (5) the highest Reynolds stresses are located within the zone of flow separation and along the shear layer; (6) high-magnitude, high-frequency quadrant-2 events ('ejections') are concentrated along the shear layer (Kelvin-Helmholtz instabilities) and dominate the contribution to the local Reynolds stress; and (7) high-magnitude, high-frequency quadrant-4 events occur bounding the separation zone, near reattachment and close to the dune crest, and are significant contributors to the local Reynolds stress at each location. These data demonstrate that the turbulence structure associated with dunes is controlled intrinsically by the formation, magnitude and downstream extent of the flow separation zone and resultant shear layer. Furthermore, the origin of dune-related macroturbulence lies in the dynamics of the shear layer rather than classical turbulent boundary layer bursting. The fluid dynamic distinction between dunes and ripples is reasoned to be linked to the velocity differential across the shear layer and hence the magnitude of the Kelvin-Helmholtz instabilities, which are both greater for dunes than ripples. These instabilities control the local flow and turbulence structure and dictate the modes of sediment entrainment and their transport rates.
1] The paper describes a numerical model for simulating sediment transport with eddyresolving 3-D models. This sediment model consists of four submodels: pickup, transport over the bed, transport in the water column and deposition, all based on a turbulent flow model using large-eddy simulation. The sediment is considered as uniform rigid spherical particles. This is usually a valid assumption for sand-bed rivers where underwater dune formation is most prominent. Under certain shear stress conditions, these particles are picked up from the bed due to an imbalance of gravity and flow forces. They either roll and slide on the bed in a sheet of sediment or separate from the bed and get suspended in the flow. Sooner or later, the suspended particles settle on the bed again. Each of these steps is modeled separately, yielding a physics-based process model for sediment transport, suitable for the simulation of bed morphodynamics. The sediment model is validated with theoretical findings such as the Rouse profile as well as with empirical relations of sediment bed load and suspended load transport. The current model shows good agreement with these theoretical and empirical relations. Moreover, the saltation mechanism is simulated, and the average saltation length, height, and velocity are found to be in good agreement with experimental results.
The present paper reports results obtained with image velocimetry to provide new insights into the two-phase nature of sediment-laden flows. The resulting two-phase flow perspective is compared with the traditional mixed-flow (or combined phase) perspective that treats sediment-laden flows essentially as flow of a single fluid. The insights are from flume experiments entailing the use of fully suspended natural sand and neutrally buoyant particles conveyed in a turbulent open channel flow of water. They confirm that suspended particles (irrespective of particle density) may affect a turbulent flow throughout its depth. Suspended particles modify flow turbulence, the main effects quantified being decreases in the bulk water velocity and in the von Kárman constant, while the flow's friction velocity remains approximately constant. Comparison of the results obtained with the two particle densities reveals differences in particle influences on water flow. In the flows conveying sand the characteristics of water and particle movement are strongly coupled, yet distinct; that is, there is a lag in the mean velocity between local water and particle movement, and intensities of water turbulence differ from intensities of particle motion turbulence. These results confirm and extend prior two-phase flow perspectives on suspended-particle transport and indicate the inaccuracies in some assumptions associated with the mixed fluid formulation of suspended-particle transport.
This report is based on the research dissertation submitted by Yarko Nifio in partial fulfillment of the requirements for the degree of Doctor of Phlosophy in Civil Engineering in the Graduate College of the University of Illinois at Urbana-Champaign in 1995. Most of the experiments on the characterization of turbulence in open channels (chapters 3 and 4) and part of the experiments on particle entrainment into suspension (chapters 7 and 8) were conducted in close collaboration with the second author, Fabikn L6pez, who also used such observations to complete h~s M.S. Thesis in Civil Engineering at the University of Illinois in 1994.
Sand settling through bedform-generated turbulence in rivers
Brendan Yuill, 1∗ (D) Yushi Wang, 1 Mead Allison, 2 Ehab Meselhe 2 and Chris Esposito 1 1 The Water Institute of the Gulf, 1110 River Road S, Suite 200, Baton Rouge, LA 70802, USA 2 Tulane University, 6823 St Charles Ave, New Orleans, LA 70118, USA
Received 28 April 2020; Revised 15 June 2020; Accepted 15 July 2020
*Correspondence to: Brendan Yuill, The Water Institute of the Gulf, 1110 River Road S, Suite 200, Baton Rouge, LA 70802, USA. E-mail: brendan.yuill@gmail.com
Fluvial bedforms, such as dunes, generate a wake with a turbulent intensity greater than the surrounding flow that may stretch distances many times the bedform length (Smith and McLean, 1977; Müller and Gyr, 1986; Lapointe, 1992; Kostaschuk and Church, 1993; Bennett and Best, 1995; Venditti and Bauer, 2005). The wake is generated from velocity shear between slow-moving flow located downstream of the bedform lee side and swifter flow passing over the bedform crest (Best, 2005). The ability for river flow to suspend bed sediment particles is dependent on the turbulent intensity of the ambient flow (Middleton and Southard, 1984; Raudkivi, 1998; van Rijn, 2007). As such, properties of particle suspension, that include the trajectory and velocity of the particles settling out of suspension, may be influenced by contact with the turbulent wake (Lapointe, 1992; Kostaschuk and Villard, 1999; Venditti and Bennett, 2000; Chang, 2004; Kwoll et al., 2013; Khosronejad and Sotiropoulos, 2014). Figure 1 shows a schematic of these phenomena and illustrates an example of how they may impact a settling sediment particle. Theoretical studies (e.g. Murray, 1970; Stout et al., 1995) suggest that isotrophic (non-directional) turbulence, on the scale that occurs in rivers, may generate a nonlinear drag effect that reduces the settling rate of sand-sized sediment particles. There has been less study
of how anisotropic (directional) turbulence, such as occurs in bedform wakes, explicitly impacts particle settling; however, processes linked to anisotropic turbulence have been shown to play a dominant role in maintaining suspension of bed material (Zedler and Street, 2001; Marchioli et al., 2006; Chang and Park, 2016; Davies and Thorne, 2016). Further, the influence of anisotropic turbulence structures, in the form of ‘sweeps’ and ‘bursts’ (Grass, 1971), on promoting the entrainment and near-bed transport of bed sediment has been well documented (McLean et al., 1994; Nelson et al., 1995; Schmeeckle, 2015).
Sediment settling has important implications for river geomorphology. The flux of suspended-bed sediment is dependent on the balance of sediment deposition due to particle settling relative to sediment entrainment from the channel bed (Raudkivi, 1998). Sediment settling in bedform fields has been shown to influence the dimensions and translation velocities of those bedforms (Hand and Bartberger, 1988; Bennett et al., 1998; Parsons and Best, 2013; Naqshband et al., 2014; Bradley and Venditti, 2019). Despite this importance, precise simulation of particle-settling processes is often neglected in geomorphic models in favour of assuming a singular particle-settling velocity, typically set as the still-water particle-settling velocity (van Rijn and Tan, 1985; Paola et al., 2011). This assumption is generally considered valid because geomorphic models are often used to simulate
FIGURE 1. Schematic of a sediment particle settling through unidirectional flow over a bedform. In the schematic, the interface between (i) the swift flow passes over the bedform and (ii) the slower flow in the separation/recirculation zone near the bedform lee side produces (iii) an area of intense velocity shear. This velocity shear produces (iv) a turbulent wake. A hypothetical sediment particle at position *a* is settling through the flow column. In the absence of turbulence, the particle will settle with a relatively uniform trajectory (*a*–*a*′) based on the pull of gravity and currents. In the presence of turbulence, the particle trajectory is affected by chaotic upward and downward pulses of velocity associated with passing eddies and, while the average vertical velocity of the eddies may be negligible, the final particle trajectory (*a*–*a*′′) and mean particle-settling rate may be significantly affected by the turbulence.
processes that operate at spatial and temporal scales much larger and longer than the hydrodynamic scales that influence particle settling (Howard, 1971; Johannesson and Parker, 1989; Murray and Paola, 1994; Nagata *et al*., 2000; Duan and Julien, 2010). However, in studies where sub-channel reach sediment transport dynamics are of interest or where settling rates are highly variable, such as near bedform-generated wakes, typical geomorphic models may lead to poor replication of the active processes. A critical step in adding precision to geomorphic models in these situations requires a better understanding of how bedform-generated turbulence might modify sediment particle settling from that expected by the still-water settling properties or by reach-averaged flow properties. Identifying if bedform-generated turbulence generally increases or decreases settling rates, and at what relative magnitude this change might occur, would be instrumental in improving this understanding (Nielsen, 1993; Wang and Maxey, 1993; Chang and Park, 2016).
In this paper, we investigate the impact of bedform-generated turbulence on sediment particles settling in river-like flow conditions using numerical modelling. We focus our investigation on how the turbulent wake generated from a single, dune-scale bedform modifies the settling velocity of medium sand. Medium sand is a primary constituent of the bed material in many medium and large sandy rivers worldwide (Molinas and Wu, 2001). Our study employs a three-dimensional (3D) computation fluid dynamics (CFD) model that resolves turbulence using detached eddy simulation (DES) and particle settling using a Lagrangian particle-tracking solver. The objectives of our study include:
Investigating how bedform-generated turbulence influences the velocity and suspended-particle fields in the local flow column;
Quantifying the simulated impact of bedform-generated turbulence on sand-sized particle-settling velocities relative to those simulated over a flat channel bed and in the absence of resolved turbulence; and
Based on the study results, discussing the implications of bedform-generated turbulence on suspended-bed sediment transport and bedform dynamics.
Methods
Overview
Our study methodology focused on the simulation of numerical model scenarios that systematically varied turbulence generation in river-like flow conditions. The objective of these scenarios was to identify the relative influence of turbulence generated from a large bedform and the channel bed on sand-sized particle settling, and to assess how this influence impacts suspended-sediment and bedform dynamics. We simulated bedform-generated turbulence by modelling the dynamic flow field over a single, fixed bedform. The bedform was based on a section of channel bed observed in the lower Mississippi River (USA) that was impacted from dredging. The dredging created a structure that had a lee-side slope with a length (20.6 m) and inclination (14°) analogous to a large natural dune (Figure 2). The lee-side slope was less than the angle of repose, which is consistent with the ‘low-angle dune’ bedform classification (Best, 2005).
Observed current-velocity profiling indicated that the engineered bedform impacted the local velocity field similar to a large natural dune (e.g. Parsons *et al*., 2005; Kwoll *et al*., 2013). Downstream of the bedform, mean velocities were significantly reduced while the mean velocity fluctuations were increased. Measurements of bed texture at the site found that the median grain-size diameter was 0.25 mm, which is typical for lower Mississippi River bars (Ramirez and Allison, 2013). Additional details about field data collected to characterize the modelled study site are available in the online Supporting Information.
The advantage of focusing the study on an engineered bedform is that a significant length of the bed upstream and downstream of the lee-side slope was relatively flat, so that the structure’s impact on the flow and sediment transport could be isolated and identified. The majority of natural dunes develop in complex fields that contain multiple dunes of irregular size, shape, and orientation (e.g. Lapointe, 1992; Parsons *et al*., 2005; Venditti and Bauer, 2005; Nittrouer *et al*., 2008; Ramirez and Allison, 2013), which makes disentangling the hydrodynamic effect of an individual bedform property extremely difficult (Wren *et al*., 2007). Past numerical and laboratory studies of bedform dynamics have adopted similar idealized approaches to simplify analyses and interpretation of results, such as by focusing on a singular structure (e.g. Nelson *et al*., 1995; Grigoriadis *et al*., 2009; Schmeeckle, 2015) or by employing idealized immobile, regularly spaced bedforms (e.g. McLean *et al*., 1994; Venditti, 2007; Kwoll *et al*., 2016).
In our model, spherical mass particles representing sediment grains were fed into suspension in unidirectionally flowing water and settled due to gravity. As the particles settled, we analysed the instantaneous and time-averaged vertical particle velocities, which we refer to as the ‘vertical particle velocity’ and the ‘particle-settling velocity’, respectively. In our study, we reference the terminal particle-settling velocity in still water, which we refer to as the ‘still-water settling velocity’. The still-water settling velocity of a particle is an intrinsic geotechnical property of an individual sediment particle and is independent of the dynamics of surrounding fluid. Our analyses focused on the settling of a medium sand-sized particle (0.25 mm in diameter), with a density of 2650 kg m<sup>−3</sup> and a terminal still-water fall velocity of 0.034 ms<sup>−1</sup>. These sediment properties pertain to the approximate median particle size and density of lower Mississippi River sand bar sediment. The still-water fall velocity approximates that predicted using the empirical formula of Dietrich (1982): 0.031 ms<sup>−1</sup>. In our study, while flow
FIGURE 2. (a) The simulated flow in this study is based on that occurring over a channel bar in the lowermost Mississippi River (USA). (b) A longitudinal transect of elevation for the study area before (undisturbed) and immediately after (initial bedform) a dredging campaign that created a large dune-scale bedform. (d) A map of the orientation of this transect. (c, e) Longitudinal and plan view morphology of a nearby natural dune field during a high river discharge (30000 m³ s⁻¹) for reference. [Colour figure can be viewed at wileyonlinelibrary.com]
and sediment particle velocity are positive net upward, settling velocity is considered positive net downward as per the convention of previous research.
Our study explicitly focused on sediment particle settling as opposed to suspended-sediment transport. Particle settling assumes the termination of movement upon particle contact with the channel bed; particles mobilized by suspended-sediment transport may continue to move upon contact with the bed, dependent on local flow properties. Suspended-sediment transport is generally modelled as the combination of particle settling and particle entrainment (van Rijn, 1984). By focusing on particle settling we can utilize a more physics-based modelling approach than if entrainment processes were also simulated. Entrainment is a complex process reliant on near-bed fluid flow, particle properties, and the composition of the surrounding bed material and it is often modelled using empirical (e.g. Engelund and Fredsøe, 1976; van Rijn, 1984; Garcia and Parker, 1991; McLean, 1992) or stochastic (e.g. Kirchner *et al.*, 1990; Papanicolaou *et al.*, 2002) approaches.
Model development
For our study, we used solvers in the OpenFOAM modeling suite (www.openfoam.org) to simulate sediment-particle movement through a turbulent, 3D open-channel flow field. Our hydrodynamic solver simulated flow dynamics using the non-hydrostatic Navier–Stokes equations for single-phase flow with constant density and viscosity. Turbulence, within the flow field, was simulated using a DES approach (i.e. the Spalart-Allmaras model; Spalart and Allmaras, 1994), which employs a one-equation mixing-length RANS (Reynolds-averaged Navier–Stokes) model to approximate boundary-layer flow near walls and a large-eddy simulation (LES) scheme away from walls. The LES scheme fully resolves turbulent flow dynamics at and above a length scale approximate to the cell dimensions of the computational grid or mesh. Filtering procedures isolate sub-grid spatial scales, where RANS modelling assumptions are also applied to predict turbulent dynamics at that (relatively small) spatial scale. We selected a DES approach for this study because the flow dynamics of interest (eddies large enough to significantly alter medium sand-sized particle settling in the context of a river channel, i.e. length scales on the order of 0.1 to 1.0 m) are located outside of the wall boundary layer in the bedform-wake zone and operate at scales much larger than the anticipated model grid-cell dimensions.
OpenFOAM is a modelling toolbox based on the finite-volume method that has become widely used in the research community because of the open-source code, easy customization, and wide array of available solvers and utilities. OpenFOAM has previously been used to accurately predict open-flow hydrodynamics in natural channels in studies focusing on horizontal recirculation in canyon rivers (Alvarez *et al.*, 2017), river channel evolution by landslides
(Zhao et al., 2017), and flow through vegetation (Chakrabarti et al., 2016), among others (e.g. Badano et al., 2012; Wang, 2013; Lai and Bandrowski, 2014).
We employed the PIMPLE solver in OpenFOAM to calculate pressure and velocity flow-field dynamics. PIMPLE employs either the PISO (pressure implicit with splitting operators) or the SIMPLE (semi-implicit method for pressure-linked equations) algorithm based on stability factors within the model during runtime that allows for adaptive timestep length. Additional numerical methods included an Euler first-order implicit scheme for time discretization, Gaussian linear for discretization of the convective terms, and Gaussian linear corrected for discretization for the Laplacian terms.
We simulated the movement of sediment suspended in flow by one-way coupling of a Lagrangian particle tracking solver (i.e. icoUncoupledKinematicParcelFoam) to the flow solver. The particle-tracking solver calculates the displacement of a prescribed number of mass particles at each time step based on the gravitational and drag forces locally affecting each particle. The particle-tracking solver was modified: (1) to calculate particle response to a time-evolving turbulent flow field and (2) to randomly fluctuate particle-injection locations over a defined depth interval. The coupled solver assigns a unique identification number to each individual particle introduced into the domain and records particle properties including velocities, age, initial injection position, and present position at each output-write interval.
In our modelling analysis, particle-settling velocity (WP) is defined as the time-averaged vertical particle velocity (UZP) over a time period of interest (T0 to T) and timestep t :
WP=T−T01∫T0TUZP(t)∂t
The dynamics of the vertical particle velocity is calculated by the particle-tracking solver from the balance of forces pertinent to a spherical object suspended in moving fluid:
∂t∂UZPmP=FZ
where mp is particle mass and FZ is the sum of the most significant forces oriented in the vertical direction, which for a sand-sized particle suspended in the flow column is the particle weight (FG) plus the vertical drag force (FDZ). Particle mass is calculated as mP=ρSπD3/6 for a particle with diameter D and density ρS. Particle weight is defined as
FG=mPg(1−ρSρI)
where g is acceleration due to gravity and ρI is the density of water. The vertical drag force exerted on a spherical particle by fluid flow can be represented by the equations
FDZ=CD8πD2ρI(UZ−UZP)∣UZ−UZP∣CD=ReP24(1+0.15ReP0.687) if ReP≤1000CD=0.44 if ReP>1000
where CD is the drag coefficient, UZ is the vertical component of the approach flow velocity, and ReP is the particle Reynold
number given by
ReP=μρID∣UZ−UZP∣
where μ is the dynamic viscosity of water. The particle relaxation time (TP) is the time it takes the particle velocity to fully respond to a change in the local vertical flow velocity; it is calculated as
TP=34ρICD∣UZ−UZP∣ρSD
and is related to the particle drag force as
FDZ=mPTPUZ−UZP
As shown in the equations above, the particle-settling velocity is dependent on the vertical drag force, which itself is dependent on the square of the vertical-velocity differential between the flow and the particle. In turbulent river flow, the vertical flow velocity is comprised of a mean vertical flow velocity (UZ) and a fluctuating component (UZ′) :
UZ=UZ+UZ′
Over distances and timescales applicable to the settling of a sediment particle in large flat-bedded rivers, UZ≈0, so for practical purposes the vertical drag force becomes dependent on the differential between the fluctuation of the vertical flow velocity and the vertical particle velocity. In turbulence research, a common metric used to characterize the intensity of the velocity fluctuations is calculated as the root-meansquare value for a time series of fluctuation values. In our study, the root-mean-square of the vertical-velocity fluctuations (UZ′RMS) is of importance and is defined as:
UZ′RMS=T−T01∫T0T[UZ′(t)]2∂t
Turbulence becomes effective at suspending particles in the flow column when UZ′RMS values approximate or exceed the particle still-water settling velocity (Middleton and Southard, 1984; Raudkivi, 1998). In rivers, undulations in the channel bed, such as bedforms, create currents with persistent nonzero vertical velocities through topographic steering; in these currents, as well as in channels with significant secondary flows, particle settling will be dependent on both UZ and UZ′ locally.
The particle-tracking solver assumed that the particle settling was dependent on the drag force exerted by the flow and the particle weight. Other forces, such as lift, were not incorporated into the modelled physics because (1) they were calculated to be an order of magnitude or less than the drag force and weight for suspended particles in our model or (2) their influential physics were not practically resolvable at our model scale. Previous studies using Lagrangian particle tracking to simulate sediment transport in rivers relied on the same simplifying assumption (e.g. Shams et al., 2002; Pasiok and StilgerSzydło, 2010; Allison et al., 2017).
Model setup
The model was set up to simulate flow and sediment transport in realistic flow conditions typical of the lower Mississippi River using the study site shown in Figure 2 as a template. The model domain was 600 m long and 20 m wide. Flow depth through the domain was variable, ranging from approximately 19 to 25 m , and based on the channel bathymetry of the simulated study site. While we focused our analyses in this study on flow and sediment transport in the downstream ( x ) and vertical ( z ) dimensions, we found that it was necessary to simulate horizontal flow (i.e. in the y dimension) to properly resolve turbulent eddy development and dissipation using the DES approach. We developed two different meshes for the model domain (Figure 3). Mesh cell size was approximately 0.33(x)×0.5(y)×0.33(z) m. Each mesh had a sub-area of cell refinement (the resolution was increased by a factor of two) designed to better resolve the flow field around the simulated bedform. The bedform was oriented so that the lee-side slope began 100 m downstream of the domain inlet. We included a zone of refinement in the mesh with a flat bed (mesh B in Figure 3) for consistency. The total number of mesh cells was 8.4 (mesh A) and 5.5 million (mesh B).
We sized the mesh cells to resolve a predominant fraction ( >80% ) of the turbulent kinetic energy generated from the bedform wake. We tested this criterion by calculating the probability distribution of the most active turbulent length scales using the integral length scale (size of the largest, most-energetic eddies) and Kolmogorov length scale (smallest effective eddy size) as reference. These turbulence scales were computed from the distributions of turbulent energy production and dissipation values predicted from an auxiliary K-epsilon RANS model of the same domain (as per the theory discussed in Umlauf and Burchard, 2003; Wang et al., 2015).
Velocity boundary conditions were set as zero gradient at the downstream boundary, ‘no-slip’ at the bottom boundary, and ‘slip’ at the surface and lateral boundaries. The inlet boundary was set as ‘mapped’, which recycled flow patterns derived from an internal location (near the outlet) to simulate a well-developed turbulent profile with a prescribed depth-averaged velocity and has been used in similar studies (e.g. Schmeeckle, 2014).
The particle-tracking model injected mass particles down a central vertical profile (extending from the water surface to a depth of -18 m ) immediately upstream of the location of the bedform crest in Figure 3a (i.e. at x=100 ). Particle injections were uniformly distributed along the vertical profile at a
prescribed rate. While sand-sized particles are typically transported vertically stratified in river flow, we chose the uniformly distributed sediment feed to simplify interpretation of the simulation results. In the case of a uniform vertical particle feed, differences in the longitudinal pattern of sediment would be predominantly a result of the properties of the flow field.
For each simulation, the model was run for 6000 s to initialize a steady-state flow field through the full model domain before the particle feed began. The total duration for the simulation of each scenario (Table 1) was set to exceed the length of time required to establish a steady-state particle deposition pattern. A steady state was assumed to be obtained when the deposition pattern became insensitive to additional simulation time as analysed at 1000 s intervals. For this analysis, once particles touched the bottom boundary (i.e. the channel bed), they became immobile and were recorded as deposited. Deposition in our model is a simplification of sediment deposition processes that occur in natural rivers where, upon contact with the channel bed, sediment particles can continue to move as bedload or become resuspended in the flow column.
The model was run on the Cypress high-performance computing (HPC) cluster at Tulane University (New Orleans, USA). Each model run was parallelized on 160 processors. Typical model run times were on the order of 5-7 days. The maximum Courant number was restricted to 0.5 ; typical model time steps ranged from 0.001 to 0.05 s for simulations employing the DES turbulence model.
Further description of model parameterization, including discretization schemes, solvers, and wall function properties, is included in the online Supporting Information.
Model scenarios
For the 3D modelling analysis, we simulated four scenarios. The text below includes a brief description of each scenario. Table 1 summarizes the key scenario properties.
Scenario 1: the reference scenario
The objective of this scenario was to investigate sedimentparticle settling over a large, simple bedform with resolved turbulence. The scenario bathymetry was set to approximate the channel bathymetry for the site shown in Figure 2b after it was modified by dredging. This bathymetry (mesh a in Figure 3) simulated a large, simplified dune-scale bedform. The lee-side slope of the bedform, located between x=100 and x=150, composed the primary morphological feature within the
FIGURE 3. Schematic of the two different model domain variations, mesh a and b. Each box shows a profile view of a domain variation; the white area is the area open to flow, the dark-grey area at the box floor represents ground and is closed to flow.
Table 1. Summary of key scenario properties
Scenario
Bed morphology a
Impact of bedform on flow
Resolved turbulence b
Bedform-generated turbulence
Background turbulence
1. Reference
bedform
yes
yes
yes
yes
2. Static velocity
bedform
yes
no
no
no
3. Flat bed
flat
no
yes
no
yes
4. Flat bed, static velocity
flat
no
no
no
no
a bedform = mesh a; flat = mesh b. b yes = employed detach eddy simulation; no = static velocity field.
domain. The mean flow velocity (1.5ms−1) through the domain inlet was set to approximate the depth-averaged velocity of a moderate discharge for the lower Mississippi River.
Scenario 2: the static-velocity field scenario
This scenario employed a static (or ‘frozen’) velocity field equal to the temporally averaged flow field of Scenario 1. This scenario simulated sediment particles settling through the same mean velocity field as Scenario 1, with no temporal variability in flow velocity. The simulation of this scenario served as a method to investigate particle settling in the absence of turbulent fluctuations in the flow field.
Scenario 3: the flat-bed scenario
The objective of this scenario was to investigate sedimentparticle settling absent of a bedform on the flow field. In this scenario, only ‘background turbulence’ (i.e. that introduced through the inlet or generated from the channel bed) was resolved. For this scenario, the relatively deep channel bed downstream of the bedform from Scenario 1 was extended upstream to fill the full domain (mesh b; see Figure 3b). The mean velocity at the inlet was set to the same value as the mean velocity in the channel downstream of the bedform from Scenario 1 (i.e. 1.18ms−1 ). Unlike the other scenarios that introduced velocity fluctuations at the inlet by mapping the velocity field to that of a downstream location with a well-developed turbulent velocity profile, in this scenario the inlet flow velocity was prescribed as a uniform value with a randomized fluctuation component. The fluctuating component of the inlet velocity was designed to replicate the same approximate turbulent intensity in the upstream channel as Scenario 1 (measured at x=50 ). This method was used to ensure that the mean ‘background turbulence’ intensity of the flow entering the model domain was approximate to that in Scenario 1.
Scenario 4: the flat-bed, static-velocity scenario
This scenario was the same as Scenario 3 except that it used a static velocity field equal to the temporally averaged flow field from Scenario 3. The objective of this scenario was to investigate sediment-particle settling in the absence of (1) turbulent fluctuations and (2) the impact of the bedform on the flow field.
Model validation
Our objective for the 3D model analyses was to simulate particle settling through realistic turbulent flow fields generally and not to precisely replicate site-specific flow hydrodynamics. Therefore, formal calibration and validation testing against observed hydrodynamic measurements lies outside the study scope. We did perform two series of sensitivity tests to
investigate how well the Spalart-Allmaras DES turbulence model simulated realistic bedform-generated wake properties in river-like flow conditions. The objective of the first series of tests was to examine how well a 3D OpenFOAM CFD model, parameterized the same as that used developed in this study, replicated the results of a robust flume study of dune-generated turbulence (Kwoll et al., 2016). The tests compared model predictions of flow velocity and turbulence with values observed in the laboratory over 30, 20, and 10∘ lee-side angle dunes. The objective of the second series of tests was to re-size the model used in the first series of tests, which had computational mesh cell dimensions on the order of 10−3m, to a larger scale suitable for simulating hydrodynamics at a real-world river scale, i.e. with mesh cell dimensions on the order of 10−1m (the same dimensions employed by the main model used in this study). The second series of tests was necessary because the simulation of turbulence is sensitive to mesh cell size; we wanted to ensure that an upscaled model which performed well in simulating hydrodynamics and turbulence at the flume scale would not perform substantially differently at the field scale.
The results of these tests indicated that the largest discrepancies between the modelled and observed hydrodynamic values were related to (1) velocity gradients very near the bed and (2) the extent of the turbulent wake. The modelled near-bed streamwise and vertical velocities had errors ranging from 5 to 20% at the mesh cell scale, with the highest errors located at the bedform-wake margins. The model tended to overestimate the area of intense turbulence in the bedform wake (from 10 to 50% ) and slightly underestimate turbulent intensity at the transition between the wake and the ambient flow column, which decreased the apparent extent of the modelled wake relative to that observed (by up to ∼25% ). The results of our mesh scaling validation exercise indicated that the range of cell sizes tested did not significantly influence the flow velocity and turbulence fields.
Interpretation of our validation results suggests that our parameterization of the hydrodynamic and turbulence model simulated flow-velocity magnitude and fluctuations in the range that occurs in real-world rivers. The primary difference between the modelled and observed flow field was that the model predicted more intense and spatially concentrated turbulence within the bedform wake. This discrepancy was likely due to the simplicity of the model domain, which could not fully replicate the variability in the flume flow. Unlike the modelled flow, the flume flow was impacted by velocity shear at the flume walls and the free surface, and by turbulence generated from mechanical flow recirculation. The impact of these phenomena would add more turbulence generally through the flow column and diffuse turbulence specifically generated in the bedform wake. The elongated flume channel would likely generate secondary currents absent in the (shorter) validation model, which would add
further discrepancies in complexity between the flume and the model flow.
Flow in natural channels would be far more complex than that simulated in the flume or in our model domains due to the presence of asymmetries in the natural channel morphology, irregular planform (e.g. channel meandering), and obstacles to flow (e.g. large sediment grains, vegetation, debris). The complex flow patterns in natural channels would likely have the net effect of further diffusing the intensity of the bedform-generated turbulence throughout the wider flow column. Therefore, our estimation of the influence of bedform-generated turbulence on particle settling in this study, which does not include this flow complexity, would likely be near the maximum strength expected in nature.
Results of the validation tests are further detailed in the online Supporting Information.
Results
Simulation of the flow field
Simulated flow over a bedform
Our model appeared to simulate realistic flow patterns and turbulence fields well; the turbulent intensities and scales of the modelled flow were on the same relative magnitude as those measured in observed river channels with similar geomorphic properties (McQuivey, 1973). In our ‘reference’ scenario, which simulated turbulent flow and particle transport over a large bedform (Figure 4), the model predicted a range of persistent flow characteristics. Depth-averaged flow velocity was reduced by 21% in the deeper channel downstream of the bedform (referred to as ‘the downstream channel’ herein) relative to the channel upstream of the bedform (referred to as ‘the upstream channel’ herein). Reach-averaged boundary shear stress was reduced by 30% on the downstream channel
bed relative to the upstream channel bed. Flow through the upstream channel maintained a relatively uniform velocity profile. The drop in bed elevation through the lee-side slope of the bedform influenced the flow similar to a naturally formed subaqueous dune (such as that documented in Parsons et al., 2005).
In the reference scenario, a significant zone of flow separation with occasional flow recirculation developed on the lee side of the bedform. This zone was typically on the order of 20 m long and intermittently extended an additional 10-20m. The largest turbulent intensities corresponded to the downstream margins of the flow separation zone ( x=140−160; see Figure 5), which was the location of the greatest velocity shear. Large turbulent eddies developed downstream of the bedform and were advected with the downstream current. These eddies typically persisted during their transport through the model domain, growing in size and becoming more amorphous with distance downstream. The eddies tended to display an advection (lift-off) angle of between 5 and 15∘.
Figure 4 b shows an example of the vertical-velocity field associated with eddies in the bedform wake. The large eddies, with characteristic dimensions ranging on the order of 100 to 101m, produced frequent and relatively steep vertical velocity gradients.
The ‘static-velocity’ scenario produced a relatively simple spatial distribution of flow velocities (Figure 6a). Flow velocities increase monotonically with distance away from the lee side of the bedform and the channel bed through the bottom half of the model domain. The steepest gradient in flow velocities occurred downstream of the bedform crest, where flow velocities transitioned from the order of 0.1 to 1.5ms−1 in a distance on the order of 5 m .
Simulated flow over a flat bed
Figures 6 b and c show example instantaneous flow fields for the ‘flat-bed’ and the ‘flat-bed, static-velocity’ scenarios,
FIGURE 4. (a) An example of the modelled instantaneous flow-velocity field extracted down the model domain centre line for the ‘reference’ scenario. (b) The vertical component of a subsection of the velocity field immediately downstream of the bedform; examples of resolved large eddy structures and the mean directionality of the bedform wake are annotated. The dashed black ovals outline the centre of vortex-like features [i.e. where the Q-criterion, a metric of the balance between fluid rotation and strain (Dubief and Delcayre, 2000) - is >0.1 ]. The white dashed box in (b) shows the approximate location of the data illustrated in Figure 6. [Colour figure can be viewed at wileyonlinelibrary.com]
**FIGURE 5.** (a) Temporally averaged vertical turbulence intensity (in terms of
UZ
-RMS) down the longitudinal centre line of the model domain for the ‘reference’ scenario. (b) Cross-section-averaged vertical-turbulence intensity by longitudinal distance for model scenarios that resolve flow turbulence. Generally,
UZ
-RMS > Ws is considered a criterion to maintain sediment suspension through diffusive mixing in alluvial channels (Raudkivi, 1998); this criterion is delineated on the plot for reference. The grey box shows the zone of high wake turbulence as interpreted from (a). Time-averaged shear velocity (u*) calculated from the ‘reference’ scenario flow field is also shown for reference. [Colour figure can be viewed at wileyonlinelibrary.com]
respectively. As stated previously, the ‘flat-bed’ scenario employed a prescribed input of uniformly distributed, random turbulence at the inlet. The level of prescribed turbulence was set to replicate the same mean intensity as that calculated in the channel upstream of the bedform in the ‘reference’ scenario. Because the ‘flat-bed’ scenario did not employ a ‘mapped’ velocity boundary condition at the inlet, a turbulent vertical-velocity profile did not fully develop until x > 100m. The flow field of the ‘flat-bed, static-velocity’ scenario displayed a logarithmic vertical-velocity profile that is typical of flow through rough-bedded open channels and was approximately uniform in the horizontal dimensions.
Predicted distributions of the vertical-velocity fluctuations
Figure 7 displays the joint distributions for the streamwise and vertical-velocity fluctuations for the ‘reference’ scenario calculated at three points distributed longitudinally, i.e. x = 150, 300, and 550. The values in the figure were all calculated at a central cross-stream location (y = 10m) and in the vertical plane with the highest mean turbulence values (z = −20m, i.e. the approximate bedform crest elevation). Turbulence data in each plot were sampled from the flow field at 20Hz for 2000s. This frequency and sampling period allowed for multiple velocity samples of a sufficient number of eddies (such as that defined in Luchik and Tiederman, 1987) to establish summary statistics that were insensitive to the addition of further data. The plots are indicative of quadrant analysis used in environmental turbulence studies that separate turbulent events into four types of turbulent event: (i) outward interactions, (ii) ejections, (iii) inward interactions, and (iv) sweeps (numerals correspond to those used in Figure 7). Typically, quadrant analysis is used to characterize turbulence generated from shear within the near-bed flow; however, here we use it to illustrate the magnitude and frequency of vertical-velocity fluctuations and their general relationship with downstream-flow fluctuations within the zone of highest turbulence intensity. In river flow, vertical-velocity fluctuations are typically considered a first-order control of suspended-sediment behaviour (Raudkivi, 1998).
In Figures 7a–c, the mean velocity fluctuation magnitudes and standard deviations decline with distance from the bedform (Table 2). The distribution of velocity fluctuations shown in Figure 7c was similar to the distribution of velocity fluctuations for the flat-bed scenario at the same location (Figure 7d), suggesting that the influence of the bedform-generated turbulence had degenerated to a similar intensity as that produced by modelled background turbulence. For the ‘reference’ scenario, ejection and sweep events were by far the most frequent types of turbulent fluctuation event. While both ejection and sweep events have been shown to positively contribute to the suspension of bed material through the positive generation of turbulence and Reynolds stresses (Bennett and Best, 1995; Nelson *et al.*, 1995; Cellino and Lemmin, 2004; Kwoll *et al.*, 2016), from a general perspective, ejection events would reduce particle-settling velocities and sweep events would increase particle-settling velocities due to their relative vertical orientations. While sweep events were somewhat more frequent than ejection events, the mean magnitudes of ejection events were greater than the mean magnitudes of sweep events.
Simulation of the suspended particles
Modelled suspended-particle fields
The simulated vertical-velocity fields appeared to have a substantial influence on the suspended-particle fields, in terms of the instantaneous vertical particle velocities (U_Zp) and particle trajectories (Figure 8). Analysis of the model results indicates that, for the vast majority of the simulation, U_Zp was equal to U_Z plus W_S (Figure 9). Differences between particle and local-flow vertical velocities were typically greatest in the most turbulent areas of flow and were attributed to the response time it took the flow current to accelerate a particle to the current velocity upon initial contact. The particle relaxation time (T_P) calculated by Equation (7) predicts that, for medium sand in the flow conditions simulated, particles reached local current velocities approximately 0.01–0.05s after initial contact with the current.
The four different scenarios simulated by the model produced substantial variations in the predicted suspended-particle fields in terms of the number of particles in suspension (relative to the particle feed at the model inlet) and in spatial distribution. Figure 10 shows the simulated suspended-particle fields at the conclusion of the model run for the four scenarios. The spatial distribution of the particles in the scenarios that resolved turbulence appears more chaotic relative to the scenarios with static-velocity fields due to mixing.
The particle mixing occurring in the turbulence-resolving scenarios maintained much more vertically uniform particle concentration profiles throughout the length of the model domain than that calculated in the scenarios without
FIGURE 7. Joint frequency distributions for streamwise (UX) and vertical (UZ) velocity fluctuations illustrated in quadrant analysis plots. Values in the quadrant corners show the percentage of the total data points present in that quadrant. (a, b, c) Data for the ‘reference’ scenario calculated at locations x=150,300,550, respectively. (d) Data for the ‘flat-bed’ scenario at x=550. [Colour figure can be viewed at wileyonlinelibrary.com]
Table 2. Summary statistic of the vertical velocity fluctuations shown in Figure 7 by quadrant
Quadrant
Scenario/location
Statistic
(i)
(ii)
(iii)
(iv)
UZz(ms−1)
UZz(ms−1)
UZz(ms−1)
UZz(ms−1)
mean
0.147
0.283
-0.169
-0.209
Ref. /x=150
Std. dev.
0.133
0.183
0.122
0.128
mean
0.062
0.091
-0.057
-0.078
Ref. /x=300
Std. dev.
0.056
0.069
0.048
0.056
mean
0.038
0.079
-0.045
-0.056
Ref. /x=550
Std. dev.
0.029
0.053
0.035
0.042
mean
0.033
0.059
-0.036
-0.044
Flat-bed /x=550
Std. dev.
0.030
0.043
0.030
0.032
turbulence (Figure 11a). This particle mixing is likely responsible for maintaining the significantly higher depthaveraged particle concentrations calculated for the ‘reference’ scenario relative to the ‘static-velocity’ scenario downstream of x=500 (Figure 11b).
Predicted particle-settling velocities
Figure 12 shows the relative impact of bedform-generated turbulence and background turbulence on the particle-settling velocities. While bedform-generated turbulence is primarily directional, background turbulence may have both significant directional (upwelling from the near-bed velocity shear) and isotropic properties (Murray, 1970). Figure 12a shows particle settling in bedform-generated turbulence and illustrates that, in the zone of intense wake turbulence ( x=100−300 ), particles were entrained in eddies developing in the velocity shear zone on the lee side of the bedform and were lifted upward through the flow column. Further downstream, the turbulence generated from the dissipating eddies kept the particles well mixed throughout the depth profile and maintained the suspension of the previously lifted particles within the upper flow column
(as suggested by the prevalence of particles with net upward trajectories, i.e. WP/WS<0 ) in the figure.
Figure 12 b shows that the particles settling through background-only turbulence (i.e. ‘the flat-bed’ scenario) were slightly more mixed through the flow column than in the absence of turbulence (i.e. the ‘flat-bed, static-velocity’ scenario), but were less mixed than when impacted by bedform-generated turbulence. Clusters of particles with similar settling velocities formed and remained relatively coherent through background-only turbulence; these clusters were less common in the ‘reference’ scenario, likely due to the more intense mixing caused by the bedform-wake turbulence.
Figure 13 illustrates the summary statics calculated for the particle-settling velocities shown in Figure 12. Figure 13a shows that the range of particle-settling velocities was much wider in the presence of bedform-generated turbulence (i.e. the ‘reference’ scenario) than in background-only turbulence (i.e. the ‘flat-bed’ scenario). For both scenarios, the distributions of particle-settling velocities stabilized through the downstream half of their domain and appeared to begin to converge to a similar, relatively small range of values. The median
FIGURE 8. A zoomed-in view of the vertical-velocity field shown in Figure 4b. The coloured arrows are flow velocity vectors for the velocity field; the black arrows show the velocity vectors for sediment particles suspended within the flow field (i.e. extracted along the same central longitudinal transect). [Colour figure can be viewed at wileyonlinelibrary.com]
FIGURE 9. Plots of the relationship between instantaneous vertical-particle velocity ( UZP ) and the surrounding vertical-flow velocity ( UZ ) field for suspended particles between x=100 and x=200 at the conclusion of the ‘reference’ scenario simulation: (a) the relationship between UZ and UZP; (b) the difference between UZ and UZP∗(UZP * is the vertical-particle velocity with the impact of the particle still-water settling velocity negated; i.e. UZP∗=UZP+WS ) by longitudinal position.
settling velocity of the ‘reference’ scenario was substantially reduced relative to the median settling velocity of the ‘staticvelocity’ scenario through the wake zone and became reduced relative to the still-water settling velocity near the downstream end of the domain. The median settling velocity for the ‘flat-bed’ scenario was generally less than the
Still-water settling velocity through the majority of the model domain.
The differences in median particle-settling velocity for the ‘reference’ and the ‘flat-bed’ scenarios relative to their counterpart scenarios that did not simulate turbulence (i.e., the ‘static-velocity’ and the ‘flat-bed, static-velocity’ scenarios, respectively) were calculated and are shown in Figure 13b. The magnitude of the difference values increased with longitudinal distance, which suggests that particle suspension became increasingly dependent on the interaction of the particles with the resolved turbulence as the particles travelled downstream. Approaching the domain outlet, turbulence resolved in the ‘flat-bed’ scenario (i.e., that resolving ‘background’ turbulence) decreased the median settling rate for suspended particles by approximately 0.01 ms<sup>−1</sup> (an ~30% reduction). The turbulence resolved in the ‘reference’ scenario (i.e., that resolving ‘background’ and ‘bedform-generated’ turbulence) decreased the median settling rate by approximately 0.025 ms<sup>−1</sup> near the domain outlet (an ~50% reduction). The reduction of particle-settling velocities due to turbulence had a significant impact on the percentage of suspended particles that were deposited on the domain bed during the associated simulations (Table 3).
Discussion
Influence of bedform-generated turbulence on particle settling
Our modeling analyses indicate that the bedform-generated wake substantially reduced the settling velocities of particles passing through it. From our analyses, we cannot precisely identify the underlying mechanics of this settling reduction; however, we can infer some of the influential processes contributing to the reduced particle settling from the modeled flow and particle fields. As we show, the vertical particle velocity was closely correlated to the vertical flow velocity. The bedform-generated turbulence does not increase mean upward oriented velocities relative to scenarios without bedform-generated turbulence; however, it may modulate
FIGURE 12. Plots of the instantaneous field of particles suspended in flow, coloured by relative settling velocity (W_{P}/W_{S}) for (a) ‘reference’ and (b) ‘flat-bed’ scenarios. The thick-dashed line delineates the upper bounds of the particle distributions for the (a) ‘static-velocity’ scenario and (b) ‘flat-bed, static-velocity’ scenario for reference. The approximate zone of the most intense particle mixing and upwelling due to wake turbulence is shown as a dashed black polygon in (a). The plots illustrate the same particle fields as shown in Figures 10a and c without reducing the number of visible particles. [Colour figure can be viewed at wileyonlinelibrary.com]
FIGURE 13. (a) Summary statistics, in box-and-whisker format, of the distributions of particle-settling velocities shown in Figure 12, binned at 50m spatial intervals in the x direction. The median particle-settling velocities estimated for the static-velocity scenarios are shown for reference; the median particle-settling velocity for the ‘flat-bed, static velocity’ scenario is approximate to the still-water settling velocity (W v ). (b) Change in median particle-settling velocity, as shown in (a), for the scenarios where turbulence is resolved relative to the static-velocity scenarios.
velocity fluctuations in a way that enhances the influence of the upward fluctuations relative to the downward fluctuations on vertical particle velocities.
Similar to observations reported in many studies of bedforms (e.g. Lapointe, 1992; Kostaschuk, 2000), our study shows that the bedform wake is primarily composed of eddy structures that are generated in the velocity shear immediately downstream of the bedform and that translate downstream and upwards, away from the bed. While these eddies do not produce a net upwards flow throughout the wake area, our modelling indicates that they do impact the vertical flow in significant ways. Figure 7 and Table 2 indicate that while downward velocity fluctuations are typically more frequent than upward velocity fluctuations, upward fluctuations are, on average, higher in magnitude. These observations appear to be aligned with the typical eddy structures qualitatively identified in Figure 4. In Figure 4, the fraction of the eddy with net downward flow is relatively large and slow moving, while the fraction composed of new upward flow was smaller and swifter, on average. We hypothesize that this asymmetry in the balance of the magnitude and frequency of upward and downward velocity fluctuations may act as a potential mechanism that results in reduced particle-settling velocities relative to the still-water particle-settling velocity; however, additional research is needed to explore this hypothesis.
Previous laboratory studies have identified instances where turbulence originating from shear at the bed (Kreplin and Eckelmann, 1979; Leeder, 1983; Wei and Willmarth, 1991) and on the lee side of bedforms (Bennett and Best, 1995) was positively skewed (net upward orientation). As observed in this study, the observed skewness resulted from shorter but higher-magnitude upward velocity fluctuations relative to downward velocity fluctuations and was capable of generating
significant net-upwards momentum transfer (Wei and Willmarth, 1991). Since Bagnold (1966) first postulated an asymmetry in vertical flow fluctuations as a driver of sediment suspension, experimental data has loosely linked its genesis to the nature of the Kelvin-Helmholtz instabilities generated in the shear layer (Bennett and Best, 1995); however, a more precise description of its origin or its role in natural river hydrodynamics and sediment transport remains largely elusive (Leeder, 1983).
Other mechanisms inherent in bedform-wake turbulence, beyond producing a skewness in the distribution of vertical-velocity fluctuations, have been identified that may result in reduced sediment particle-settling rates. Studies on the motion of sand particles in turbulent flow (Stommel, 1949; Tooby et al., 1977; Nielsen, 1984) have shown that sand-sized particles may become trapped within the vortex of the eddies composing the bedform wake. Logically, the particles trapped within the rotating flow volume in the eddy vortex would be swept preferentially in the mean direction of the eddy and, as the eddies typically have a substantial mean upward trajectory, the downward settling movement of the particles would be reversed or significantly reduced. This ‘eddy-trapping’ process has been used as an explanation for the high sediment concentrations associated with turbulent boils observed in rivers (Bijker et al., 1976; Shugar et al., 2010), although the mechanics of the process have not been well explored in river-like flow conditions.
Schmeeckle et al. (1999) and Venditti and Bennett (2000) describe mechanisms closely related to eddy trapping that occur during eddy genesis over bedforms. Using a combination of particle-imaging velocimetry and numerical modelling, Schmeeckle et al. (1999) describe two mechanisms that
Table 3. Predicted particle deposition values upon the model domain bed for each scenario
a Relative to the ‘static-vel.’ scenario. b Relative to the ‘flat-bed, static-vel.’ scenario. ↩︎
contribute to the net upward flux of near-bed suspended sediment: (1) an expansion of the near-bed shear layer at a bedform crest preceding eddy development that sucks near-bed sediment upwards and (2) the inclusion of flow with relatively high sediment concentrations sourced from within the recirculation zone on the bedform lee side into the eddy structure during the eddy development and shedding process. In a laboratory experiment employing fixed acoustic Doppler velocimetry and optical turbidity sensors, Venditti and Bennett (2000) generally confirmed the results of Schmeeckle et al. (1999) and found that near-bed sediment launched higher into the flow column by bedform-generated turbulence accounted for nearly all of the sediment suspension occurring during their experiment.
Bradley et al. (2013) provide field evidence that dunes may promote locally high sediment transport concentrations without the production of a significant turbulent wake. In a study of low-angled dunes in the estuarine reach of the Fraser River, they found that vertical, topographically induced currents generated over the stoss-side slope of a dune were responsible for lifting significant fractions of near-bed sediment into the flow column and served as the source for up to 69% of the total suspended-sediment transport passing over the dune. In a similar field study of an estuarine reach of the Elba River, Kwoll et al. (2013) found that tides heavily modulated the influence of the turbulent wake on suspended-sediment transport. They report that, during tidally induced periods of slack water, low river velocities passing over bedforms did not generate adequate turbulence to influence suspended-sediment behaviour. As tides shifted and near-bed flow velocities accelerated, turbulence structures generated in the lee of bedforms became increasingly important drivers of total suspended-sediment transport flux. In the final stage, when the tidal conditions produced the maximum channel flow velocities, the influence of the bedform-generated turbulence on the suspended-sediment flux deteriorated as the influence of bed stress became the dominate driver.
Our numerical results show that particles are preferentially located within eddy structures and that the eddies are efficient at lifting near-bed sediment particles higher into the flow column. This phenomenon has been documented in the field for decades (e.g. Kostaschuk and Church, 1993); however, high-resolution illustration of the interaction between turbulence and suspended particles, as presented in our study, is rare. Figure 12 shows that in the downstream model domain, the upper flow column is predominantly populated with sediment particles lifted upwards from a near-bed position by the wake turbulence, evidenced by the preponderance of particles with negative mean settling velocities in that area.
As referenced in the Introduction, there has been considerable research on the effect of nonlinear drag force on particle settling. While this effect, which can impact sediment settling in isotropic (non-directional) turbulence, has primarily been investigated in well-constrained flow conditions numerically or in the laboratory (Nielsen, 1993; Brucato et al., 1998; Fung, 1998; Fornari et al., 2016), it has been hypothesized to have a significant impact on suspended sediment in rivers (Murray, 1970). To approximate the general magnitude of the nonlinear drag effect on particle settling in the hydrodynamic conditions simulated in this study, we performed a simple one-dimensional (1D) modelling experiment (fully described in the online Supporting Information). In our 1D experiment, we calculated particle settling using the same equations of particle movement as in our 3D model [Equations (1)-(10)] but with simplified flow velocity: horizontal flow was set to zero, while vertical flow velocity was approximated as a sine wave with zero mean value. We simulated a range of turbulent intensities by varying the amplitude and frequency of the sine wave
in a manner similar to Murray (1970) and Stout et al. (1995). We found that for turbulence intensities in the range observed in the lower Mississippi River (McQuivey, 1973), the nonlinear drag effect may significantly reduce mean particle-settling velocities up to approximately 10%. However, the largest realistic turbulence intensities tested did not reduce particle-settling rates to the extent observed in our 3D modelling results, which suggests that mechanisms related to anisotropic turbulence have a far greater influence on sediment settling through bedform-generated turbulence.
Our study simulated turbulence generated from a bedform with a 14∘ lee-side slope angle. While we did not perform statistical tests to identify the frequency of the instances of predicted flow separation, qualitative observations suggest that a significant fraction of the flow separation occurring above the bedform lee-side slope was permanent. Field and laboratory observations (e.g. Kostaschuk and Villard, 1996; Best and Kostaschuk, 2002; Lefebvre et al., 2014; Kwoll et al., 2016) of flow over dunes suggest that permanent flow separation typically only occurs over the lee-side slope near the angle of repose (i.e. ∼30∘ ), although other numerical modelling studies (e.g. Paarlberg et al., 2009; Lefebvre and Winter, 2016) have predicted that permanent flow separation may occur from dunes with lee-side slope angles between 10 and 20∘. Research on dunes with variable lee-side slope angles typically suggests that smaller angles produce systematically smaller and more intermittent instances of separated flow (Best, 2005; Kwoll et al., 2016).
Our numerical model simulated 3D flow, however, for simplicity, the bedform morphology incorporated into the model domain did not vary in the cross-stream dimension. The influence of 3D properties of dunes and dune fields on reach-scale flow, sediment transport, and channel morphology has been well documented in the field (e.g. Dietrich and Smith, 1984; Parsons et al., 2005; Herbert and Alexander, 2018), laboratory (e.g. Venditti, 2007), and by numerical modelling (e.g. Johns and Xing, 1993; Omidyeganeh and Piomelli, 2013; Chen et al., 2015). Of particular relevance to our study, Lefebvre (2019) used numerical modelling to show that the lateral orientation of the lee-side slope relative to the mean direction of the flow passing over the bedform is a primary control of the lee-side flow separation and turbulent wake production. Lefebvre (2019) found that as the lee-side slope became less perpendicular to the mean flow direction, flow separation was reduced. The exclusion of bedform three-dimensionality in our model should be considered when extrapolating our results to natural rivers.
This study focused on how bedform-generated turbulence impacts the settling of medium sand-sized particles. The settling behaviour of different particle size fractions may be substantially different than that identified by our study results. Medium sand settling was strongly sensitive to the range of bedform-generated turbulence simulated in our study. This sensitivity was likely promoted by the fact that, as medium sand particles settled from the outer flow column into the turbulent wake zone, the ratio of their WS to the local turbulent intensity (UZ′/kMS), which may serve as a criterion for transport mode (Kaudkivi, 1998), became increasingly supportive of particle suspension (shifting from WS/UZ′/kMS>1 to WS/UZ′/kMS<1 ). Settling of particle-size fractions that do not experience a significant shift in their WS/UZ′/kMS ratio upon contact with the wake zone, either maintaining consistent strong suspension (WS/UZ′/kMS<<1) or a lack of suspension (WS/UZ′/kMS>>1), would be relatively insensitive to bedform-generated turbulence. Based on the range of turbulence intensities predicted for the bedform wakes in this study, the settling behaviour of particles smaller than fine sand ( WS for fine sand =0.01ms−1 )
or larger than very coarse sand (WS for coarse sand =0.15ms−1) would not likely be significantly impacted.
Implications for suspended-sediment concentrations and bedform dynamics
Our study illuminates how turbulence contributes to a reduction in sediment-settling rates, which may have a substantial effect on suspended-sediment transport and bedform dynamics. A general reduction in settling rates increases the time a particle of bed sediment will spend suspended in the flow column and, if particle entrainment processes are negligibly affected by the reduced settling (e.g. if there is no corresponding sediment-supply limitations or turbulence dampening) (VanSickle and Beschta, 1983; Sheng and Villaret, 1989), will result in higher suspended-sediment concentrations. While previous studies focusing on bedform-generated turbulence often identify higher concentrations of suspended sediment in the bedform-wake area (Lapointe, 1992; Kostaschuk and Church, 1993; Venditti and Bennett, 2000), the effect of bedforms on reach-scale suspended-sediment concentrations is less well understood as bedforms modify sediment transport by additional (sometimes offsetting) mechanisms, such as by increasing flow resistance (Smith and McLean, 1977; Naqshband et al., 2014).
Increasing suspended-sediment transport over subaqueous dunes has typically been shown to flatten the dune morphology (Bridge and Best, 1988; Bennett et al., 1998; Kostaschuk and Best, 2005; Hendershot et al., 2016; Bradley and Venditti, 2019). This flattening may result from a reduction in the angle of the lee-side slopes and from a decreased relative contribution of bedload transport, which predominantly drives bedform growth and steepening (Simons et al., 1965; Carling et al., 2000; Jerolmack and Mohrig, 2005) to the overall sediment flux. When increased suspended-sediment transport is also associated with a significant increase in sediment transport capacity, apparent dune flattening may result from the transition to an upper-stage plane bed (Simons and Richardson, 1961). There are some contrasting observations from natural channels that identified a positive relationship between suspended-sediment transport and dune height (e.g. Allen, 1978; Gabel, 1993; Amsler and Schreider, 1999), but in those instances the increase in dune height resulted from amalgamation of multiple dunes (Best, 2005; Reesink and Bridge, 2007).
In the event that dune flattening is caused by reduced slope angles, it is generally assumed to produce less flow acceleration over the stoss-side slope (Nelson et al., 1993) and a smaller zone of flow separation on the lee side of the dune (Best and Kostaschuk, 2002; Kwoll et al., 2016), which results in a less intense turbulent wake. As suggested by this study, reduced wake intensity generally promotes faster particle settling and could potentially lead to decreased suspended-sediment transport rates within the bedform-wake zone. However, the impact of dune flattening on reach-scale suspended-sediment transport is less certain. Lefebvre and Winter (2016) show that the hydraulic roughness produced by bedforms is positively related to the lee-side slope angle and dune height relative to flow depth. Reduced roughness results in increased flow velocities that would promote increased suspended transport rates generally (Raudkivi, 1998).
A relatively explicit result of this study is that suspendedsediment particles passing over a bedform crest will, on average, contact the channel bed further downstream under the influence of turbulence than without it. In the context of a dune field, by increasing the area over which particles
passing over a dune will contact the bed, turbulence ‘diffuses’ the contribution of those particles to a larger number of downstream dunes. In cases where the deposition of suspended sediment plays an integral part in the dune dynamics, this increase in mean settling trajectory length could ultimately lead to the diffusion of dune properties (e.g. flattening or increased wave length) and, as particles moving in suspension have higher average velocities than particles moving in contact with the bed, higher bedform translation velocities (Hand and Bartberger, 1988; Prent and Hickin, 2001; McElroy and Mohrig, 2009; Parsons and Best, 2013; Naqshband et al., 2014). In cases where suspended sediment is not a significant driver of dune dynamics, increased trajectory lengths would lead to higher instances of particle ‘bypassing’ for individual dunes (Mohrig and Smith, 1996; Naqshband et al., 2014). In the model of dune evolution postulated by Mohrig and Smith (1996), particles suspended over a dune crest must interact with the stoss-side slope of the downstream dune if that particle is going to contribute to the net translation of dunes within a field; particles bypassing the stoss-side slope contribute to the fraction of the sediment load not represented by the bedform sediment flux. Given the assumptions of that model, by increasing the settling trajectory length, bedform-generated turbulence decreases the overall flux of sediment transported in bedforms. It is logical to assume that, in cases where particles bypass one dune to contribute to another dune further downstream, the bypass would lead to decreased symmetry between neighbouring dunes rather than an overall reduction in bedform sediment flux.
Implications for suspended-sediment transport and morphological modelling
A common method to model suspended-sediment transport is to incorporate the Rouse number as a criterion for particle suspension. While initially developed to predict equilibrium sediment concentration profiles under uniform flow conditions over a flat-plane bed, models reliant on a Rouse number-type approach have been used to predict suspended-sediment transport in a range of dynamic hydrodynamic environments (Gelfenbaum and Smith, 1986). Our study provides additional evidence that the Rouse number is not necessarily applicable in the proximity of bedforms (Akins et al., 1989; Bennett et al., 1998; Schmeeckle et al., 1999). The Rouse number relies on the near-bed shear velocity (u∗) as an estimate of the net-upwards vertical force balancing the downward pull of gravity on a particle (McLean, 1992; Leeder et al., 2005); however, as shown in our study, the typically strong correlation between u∗ and USsgAB (Raudkivi, 1998) breaks down under the influence of bedform-wake turbulence (e.g. Figure 5). This lack of correlation derives from the fact that much of the bedform wake is generated from shear at the flow separation zone on the lee side of the bedform rather than shear at the bed, which may have a significantly different velocity distribution (Nelson et al., 1995).
Due to the substantial effect of bedform-generated turbulence on sediment settling, morphodynamic models that disregard this effect may over-predict sediment deposition in the area downstream of the bedform and misrepresent sediment transport rates generally. As observed in our study, the presence of turbulent structures that significantly influence sediment settling and suspension may become unidentifiable in the mean flow field (Chang et al., 2011). Suspended-sediment models based on temporal-averaged hydrodynamics and gradient
diffusion would not resolve these turbulent structures or the relatively high loads of sediment transported within them. Further, models that do not fully resolve the key structures of the wake turbulence, such as the eddy dynamics (e.g. those that use turbulence statistics to incorporate the mean impact of turbulence, such as RANS-based models), may not accurately predict sediment deposition in the wake zone (Keylock et al., 2014). This is because the properties of the discrete turbulence fluctuations, in terms of the fluctuation magnitudes and durations, have been shown to be substantially better correlated to the resultant sediment transport than the temporally averaged fluctuation values due to the highly nonlinear positive relationship between flow velocity and sediment flux (McLean et al., 1994; Nelson et al., 1995; Hurther and Lemmin, 2003; Bhaganagar and Hsu, 2009; Lelouvetel et al., 2009; Schmeeckle, 2015). Recent research (e.g. Chang and Scotti, 2003; Tominaga and Stathopoulos, 2011; Keylock et al., 2012; Alvarez et al., 2017) has shown how turbulent-resolving models, such as those employing LES, more accurately predict flow and particulate transport than models reliant on turbulent statistics.
In geomorphic models of river systems, bedforms are often assumed to reduce suspended-sediment transport by adding roughness to the channel bed and decelerating the overall flow velocity (van Rijn, 1984; Garcia and Parker, 1991), despite a growing breadth of field evidence that suggests this assumption may not be universally applicable (Rood and Hickin, 1989; Kostaschuk and Church, 1993; Venditti and Bennett, 2000; Kostaschuk et al., 2009; Bradley et al., 2013). As suggested by our study, it is probable that due to the reduced settling in the bedform-wake zone, bedforms increase suspended sediment-transport rates locally in certain configurations. Studies of submerged vegetation (e.g. Yang, 2007; Ortiz et al., 2013; Yager and Schmeeckle, 2013), which is also often regarded as a roughness element similar to dunes (Vargas-Luna et al., 2015), have shown that it can promote sediment transport through the persistent generation of turbulence dependent on stem morphology and density (Nepf, 1999; Yang and Nepf, 2018).
Conclusions
Our study objective was to quantify bedform-generated turbulence influences on medium sand-sized particle settling in river-like flow conditions and assess how that influence might impact suspended-sediment and bedform dynamics. We used a numerical model to simulate 3D flow and suspended-particle fields to investigate sediment settling through the wake of a large, dune-scale bedform. To isolate the effect of turbulence on sediment settling, we simulated scenarios with and without temporally fluctuating flow. To differentiate the effect of the bedform wake on particle settling relative to that of background turbulence alone, we simulated scenarios with and without a bedform present on the channel bed.
On average, the presence of bedform-generated turbulence increased turbulence intensity within the model domain and generally reduced particle-settling velocities. Local values of particle-settling velocity and the settling trajectory of individual particles were significantly influenced by anisotropic turbulent structures resolved within the flow-velocity field. Over the model domain, bedform-generated plus background turbulence reduced the settling velocity of suspended particles by approximately 50% relative to that estimated in flow without turbulent fluctuations. This reduction in settling velocity resulted in 24% fewer particles becoming deposited within the model domain. Background turbulence alone was
calculated to have decreased settling velocity by up to 30% and decreased deposition within the domain by 5%. Our study examined a bedform with a relatively low lee-side slope angle (14∘). Bedforms with larger angles near the angle of repose (∼30∘) would likely generate more intense wakes with a greater impact on particle settling.
The vertical-particle velocity was only significantly different than the vertical-flow velocity for very short time periods after high-magnitude fluctuations in the vertical-flow velocity. The mechanism by which turbulence reduced particle settling was attributed to the production of positively skewed vertical-velocity distributions. These distributions appeared to vary in space and were related to the mean turbulence intensity of the flow field.
The interplay between bedforms, the generation of turbulence, and sediment transport is complex, and the relative impact of the related individual processes is difficult to detangle from one another. Numerical modelling studies, such as this study, that can simulate synthetic scenarios designed to isolate the impact of processes and properties that cannot be isolated in the real world, in a realistic physics-driven manner, may be of great use to improve our scientific understanding. To simplify our study methodology, the basis of our analysis was a single, simplified bedform and we limited our investigation to the particle-settling process and explicitly did not simulate sediment-particle entrainment.
Acknowledgements-This study was funded by The Water Institute of the Gulf Science and Engineering Plan and the Louisiana Coastal Protection and Restoration Authority. Diana Di Leonardo helped improve the clarity of this manuscript. Luca Solari provided editorial support of a previous version of this manuscript that improved its quality.
Data Availability Statement
The data that support the findings of this study are available from the corresponding author upon reasonable request.
Conflict of Interest
The authors declare no conflict of interest.
References
Allen JRL. 1978. Polymodal dune assemblages: an interpretation in terms of dune creation-destruction in periodic flows. Sedimentary Geology 20: 17-28.
Allison MA, Yuill BT, Meselhe EA, Marsh JK, Kolker AS, Ameen AD. 2017. Observational and numerical particle tracking to examine sediment dynamics in a Mississippi River delta diversion. Estuarine, Coastal and Shelf Science 194: 97-108. https://doi.org/10.1016/j. ecss.2017.06.004
Alvarez LV, Schmeeckle MW, Grams PE. 2017. A detached eddy simulation model for the study of lateral separation zones along a large canyon-bound river. Journal of Geophysical Research: Earth Surface 122: 25−49.
Amiler ML, Schreider MI. 1999. Dune height prediction at floods in the Paraná River, Argentina. In River Sedimentation: Theory and Applications, Jayewardena AW, Lee JHW, Wang ZY (eds). A.A. Balkema: Rotterdam; 615-620.
Atkins R, Soulsby RL, Waters CB, Oliver N. 1989. Field measurements of sediment suspension above bedforms in a sandy estuary. Technical Report. Hydraulics Research: Wallingford.
Badano ND, Sabarots Gerbec M, Re M, Menéndez AN. 2012. A coupled hydrosedimentologic model to assess the advance of the Parana River Delta Front. In Proceedings of the Sixth International Conference on Fluvial Hydraulics, San Jose, Costa Rica.
Bagnold RA. 1966. An Approach to the Sediment Transport Problem from General Physics. US Government Printing Office: Washington, D.C.
Bennett SJ, Best JL. 1995. Mean flow and turbulence structure over fixed, two-dimensional dunes: implications for sediment transport and bedform stability. Sedimentology 42: 491-513. https://doi.org/ 10.1111/j.1365-3091.1995.tb00386.x
Bennett SJ, Bridge JS, Best JL. 1998. Fluid and sediment dynamics of upper stage plane beds. Journal of Geophysical Research: Oceans 103: 1239-1274. https://doi.org/10.1029/97JC02764
Best J. 2005. The fluid dynamics of river dunes: a review and some future research directions. Journal of Geophysical Research: Earth Surface 110: 1-21. https://doi.org/10.1029/2004JF000218
Best J, Kostaschuk R. 2002. An experimental study of turbulent flow over a low-angle dune. Journal of Geophysical Research: Oceans 107: 18-1-18-19.
Bhaganagar K, Hsu T-J. 2009. Direct numerical simulations of flow over two-dimensional and three-dimensional ripples and implication to sediment transport: steady flow. Coastal Engineering 56: 320-331. https://doi.org/10.1016/j.coastaleng.2008.09.010
Bijker EW, Van Hijum E, Vellinga P. 1976. Sand transport by waves. In Coastal Engineering Proceedings.
Bradley RW, Venditti JG. 2019. The growth of dunes in rivers. Journal of Geophysical Research: Earth Surface 124: 548-566.
Bradley RW, Venditti JG, Kostaschuk RA, Church M, Hendershot M, Allison MA. 2013. Flow and sediment suspension events over low-angle dunes: Fraser Estuary, Canada. Journal of Geophysical Research: Earth Surface 118: 1693-1709.
Bridge JS, Best JL. 1988. Flow, sediment transport and bedform dynamics over the transition from dunes to upper-stage plane beds: implications for the formation of planar laminae. Sedimentology 35: 753−763.
Brucato A, Grisafi F, Montante G. 1998. Particle drag coefficients in turbulent fluids. Chemical Engineering Science 53: 3295-3314.
Carling PA, Williams JJ, Golz E, Kelsey AD. 2000. The morphodynamics of fluvial sand dunes in the River Rhine, near Mainz, Germany. II. Hydrodynamics and sediment transport. Sedimentology 47(1), 253-278.
Cellino M, Lemmin U. 2004. Influence of coherent flow structures on the dynamics of suspended sediment transport in open-channel flow. Journal of Hydraulic Engineering 130: 1077-1088.
Chakrabarti A, Chen Q, Smith HD, Liu D. 2016. Large eddy simulation of unidirectional and wave flows through vegetation. Journal of Engineering Mechanics 142: 1-8, 04016048.
Chang YS. 2004. Suspended sediment and hydrodynamics above mildly sloped long wave ripples. Journal of Geophysical Research 109: 1-16. https://doi.org/10.1029/2003JC001900
Chang YS, Park Y-G. 2016. Suspension of sediment particles over a ripple due to turbulent convection under unsteady flow conditions. Ocean Science Journal 51: 127-135. https://doi.org/10.1007/ s12601-016-0011-2
Chang Y, Scotti A. 2003. Entrainment and suspension of sediments into a turbulent flow over ripples. Journal of Turbulence 4: 1-22. https:// doi.org/10.1088/1468-5248/4/1/019
Chang WY, Constantinescu G, Tsai WF, Lien HC. 2011. Coherent structure dynamics and sediment erosion mechanisms around an in-stream rectangular cylinder at low and moderate angles of attack. Water Resources Research 47(W12532), 1-16. https://doi.org/ 10.1029/2011WR010586.
Chen X, Cardenas MB, Chen L. 2015. Three-dimensional versus two-dimensional bed form-induced hyporheic exchange. Water Resources Research 51: 2923-2936.
Davies AG, Thorne PD. 2016. On the suspension of graded sediment by waves above ripples: inferences of convective and diffusive processes. Continental Shelf Research 112: 46-67. https://doi.org/ 10.1016/j.csr.2015.10.006
Dietrich WE. 1982. Settling velocity of natural particles. Water Resources Research 18: 1615-1626. https://doi.org/10.1029/ WR018006p01615
Dietrich WE, Smith JD. 1984. Bed load transport in a river meander. Water Resources Research 20: 1355-1380.
Duan JG, Julien PY. 2010. Numerical simulation of meandering evolution. Journal of Hydrology 391: 34-46.
Dubief Y, Delcayre F. 2000. On coherent-vortex identification in turbulence. Journal of Turbulence 1: 1-22. https://doi.org/10.1088/1468$5248 / 1 / 1 / 011$
Engelund F, Fredsøe J. 1976. A sediment transport model for straight alluvial channels. Hydrology Research 7: 293-306.
Fornari W, Picano F, Brandt L. 2016. Sedimentation of finite-size spheres in quiescent and turbulent environments. Journal of Fluid Mechanics 788: 640-669. https://doi.org/10.1017/jfm. 2015.698
Fung JCH. 1998. Effect of nonlinear drag on the settling velocity of particles in homogeneous isotropic turbulence. Journal of Geophysical Research: Oceans 103: 27905-27917. https://doi.org/10.1029/ 98JC02822
Gabel SL. 1993. Geometry and kinematics of dunes during steady and unsteady flows in the Calamus River, Nebraska, USA. Sedimentology 40: 237-269.
Garcia M, Parker G. 1991. Entrainment of bed sediment into suspension. Journal of Hydraulic Engineering 117: 414-435.
Gelfenbaum G, Smith JD. 1986. Experimental evaluation of a generalized suspended-sediment transport theory. In Shelf Sands and Sandstones, Knight RJ, McLean JR (eds). Canadian Society of Petroleum Geologists: Calgary; 133-134.
Grass AJ. 1971. Structural features of turbulent flow over smooth and rough boundaries. Journal of Fluid Mechanics 50: 233-255.
Grigoriadis DGE, Balaras E, Dimas AA. 2009. Large-eddy simulations of unidirectional water flow over dunes. Journal of Geophysical Research 114(F02022), 1-19. https://doi.org/10.1029/2008JF001014
Hand BM, Bartberger CE. 1988. Leeside sediment fallout patterns and the stability of angular bedforms. Journal of Sedimentary Research 58: 33-43. https://doi.org/10.1306/212F8D05-2B24-11D78648000102C1865D
Hendershot ML, Venditti JG, Bradley RW, Kostaschuk RA, Church M, Allison MA. 2016. Response of low-angle dunes to variable flow. Sedimentology 63: 743-760.
Herbert CM, Alexander J. 2018. Bottomset architecture formed in the troughs of dunes and unit bars. Journal of Sedimentary Research 88: 522−553.
Howard AD. 1971. Simulation of stream networks by headword growth and branching. Geographical Analysis 3: 29-50.
Hurther D, Lemmin U. 2003. Turbulent particle flux and momentum flux statistics in suspension flow. Water Resources Research 39(5), 1139-1150. https://doi.org/10.1029/2001WR001113.
Jerolmack DJ, Mohrig D. 2005. A unified model for subaqueous bed form dynamics. Water Resources Research 41(W12421), 1-10. https://doi.org/10.1029/2005WR004329.
Johannesson H, Parker G. 1989. Linear theory of river meanders. River Meandering 12: 181-213.
Johns B, Xing J. 1993. Three-dimensional modelling of the free surface turbulent flow of water over a bedform. Continental Shelf Research 13: 705-721.
Keylock CJ, Constantinescu G, Hardy RJ. 2012. The application of computational fluid dynamics to natural river channels: eddy resolving versus mean flow approaches. Geomorphology 179: 1-20. https:// doi.org/10.1016/j.geomorph.2012.09.006
Keylock CJ, Lane SN, Richards KS. 2014. Quadrant/octant sequencing and the role of coherent structures in bed load sediment entrainment. Journal of Geophysical Research: Earth Surface 119: 264-286. https:// doi.org/10.1002/2012JF002698
Khosronejad A, Sotiropoulos F. 2014. Numerical simulation of sand waves in a turbulent open channel flow. Journal of Fluid Mechanics 753: 150-216. https://doi.org/10.1017/jfm. 2014.335
Kirchner JW, Dietrich WE, Iseya F, Ikeda H. 1990. The variability of critical shear stress, friction angle, and grain protrusion in water-worked sediments. Sedimentology 37: 647-672.
Kostaschuk R. 2000. A field study of turbulence and sediment dynamics over subaqueous dunes with flow separation. Sedimentology 47: 519−531.
Kostaschuk R, Best J. 2005. Response of sand dunes to variations in tidal flow: Fraser Estuary. Canada. Journal of Geophysical Research: Earth Surface 110(F04504), 1-10. https://doi.org/10.1029/ 2004JF000176
Kostaschuk RA, Church MA. 1993. Macroturbulence generated by dunes: Fraser River, Canada. Sedimentary Geology 85: 25-37. https://doi.org/10.1016/0037-0738(93)90073-E
Kostaschuk R, Villard P. 1996. Flow and sediment transport over large subaqueous dunes: Fraser River, Canada. Sedimentology 43: 849-863.
Kostaschuk RA, Villard PV. 1999. Turbulent sand suspension over dunes. In Fluvial Sedimentology VI, Vol. 28: 3-14.
Kostaschuk R, Shugar D, Best J, Parsons D, Lane S, Hardy R, Orfeo O. 2009. Suspended sediment transport and deposition over a dune: Rio Paraná, Argentina. Earth Surface Processes and Landforms 34: 1605−1611.
Kreplin H-P, Eckelmann H. 1979. Behavior of the three fluctuating velocity components in the wall region of a turbulent channel flow. The Physics of Fluids 22: 1233-1239.
Kwoll E, Winter C, Becker M. 2013. Intermittent suspension and transport of fine sediment over natural tidal bedforms. In Coherent Flow Structures at Earth’s Surface, Venditti JG, Best JL, Church M, Hardy RJ (eds). Wiley-Blackwell: Oxford; 231-242.
Kwoll E, Venditti JG, Bradley RW, Winter C. 2016. Flow structure and resistance over subaqueous high- and low-angle dunes. Journal of Geophysical Research: Earth Surface 121: 545-564. https://doi.org/ 10.1002/2015JF003637
Lai YG, Bandrowski DJ. 2014. Large wood flow hydraulics: a 3D modelling approach. In Proceedings of Bold Visions for Environmental Modelling, San Diego, CA.
Lapointe M. 1992. Burst-like sediment suspension events in a sand bed river. Earth Surface Processes and Landforms 17: 253-270. https:// doi.org/10.1002/esp. 3290170305
Leeder MR. 1983. On the dynamics of sediment suspension by residual Reynolds stresses-conformation of Bagnold’s theory. Sedimentology 30: 485-491.
Leeder MR, Gray TE, Alexander J. 2005. Sediment suspension dynamics and a new criterion for the maintenance of turbulent suspensions. Sedimentology 52: 683-691.
Lefebvre A. 2019. Three-dimensional flow above river bedforms: insights from numerical modeling of a natural dune field (Rio Paraná, Argentina). Journal of Geophysical Research: Earth Surface 124: 2241-2264.
Lefebvre A, Winter C. 2016. Predicting bed form roughness: the influence of lee side angle. Geo-Marine Letters 36: 121-133.
Lefebvre A, Paarlberg AJ, Winter C. 2014. Flow separation and shear stress over angle-of-repose bed forms: a numerical investigation. Water Resources Research 50: 986-1005. https://doi.org/10.1002/ 2013WR014587
Lelouvetel J, Bigillon F, Doppler D, Vinkovic I, Champagne J-Y. 2009. Experimental investigation of ejections and sweeps involved in particle suspension. Water Resources Research 45(W02416), 1-15. https://doi.org/10.1029/2007WR006520
Luchik TS, Tiederman WG. 1987. Timescale and structure of ejections and bursts in turbulent channel flows. Journal of Fluid Mechanics 174: 529-552.
Marchioli C, Armenio V, Salvetti MV, Soldati A. 2006. Mechanisms for deposition and resuspension of heavy particles in turbulent flow over wavy interfaces. Physics of Fluids 18: 1-16, 025102. https://doi.org/ 10.1063/1.2166453
McElroy B, Mohrig D. 2009. Nature of deformation of sandy bed forms. Journal of Geophysical Research 114(F00A04), 1-13. https://doi.org/ 10.1029/2008JF001220
McLean SR. 1992. On the calculation of suspended load for noncohesive sediments. Journal of Geophysical Research: Oceans 97: 5759-5770.
McLean SR, Nelson JM, Wolfe SR. 1994. Turbulence structure over two-dimensional bed forms: implications for sediment transport. Journal of Geophysical Research 99(C6), 12729-12747. https://doi. org/10.1029/94JC00571
McQuivey RS. 1973. Summary of Turbulence Data from Rivers, Conveyance Channels, and Laboratory Flumes: Turbulence in Water. US Government Printing Office: Washington, D.C.
Middleton GV, Southard JB. 1984. Mechanics of sediment movement. Short Course No. 3. Society of Economic Paleontologists and Mineralogists: Providence, RI.
Mohrig D, Smith JD. 1996. Predicting the migration rates of subaqueous dunes. Water Resources Research 32: 3207-3217. https://doi. org/10.1029/96WR01129
Molinas A, Wu B. 2001. Transport of sediment in large sand-bed rivers. Journal of Hydraulic Research 39: 135-146.
Müller A, Gyr A. 1986. On the vortex formation in the mixing layer behind dunes. Journal of Hydraulic Research 24: 359-375.
Murray SP. 1970. Settling velocities and vertical diffusion of particles in turbulent water. Journal of Geophysical Research 75: 1647-1654. https://doi.org/10.1029/JC075009p01647
Murray AB, Paola C. 1994. A cellular model of braided rivers. Nature 371: 54-57.
Nagata N, Hosoda T, Muramoto Y. 2000. Numerical analysis of river channel processes with bank erosion. Journal of Hydraulic Engineering 126: 243-252.
Naqshband S, Ribberink JS, Hurther D, Hulscher SJ. 2014. Bed load and suspended load contributions to migrating sand dunes in equilibrium. Journal of Geophysical Research: Earth Surface 119: 1043-1063.
Nelson JM, McLean SR, Wolfe SR. 1993. Mean flow and turbulence fields over two-dimensional bed forms. Water Resources Research 29: 3935-3953. https://doi.org/10.1029/93WR01932
Nelson JM, Shreve RL, McLean SR, Drake TG. 1995. Role of near-bed turbulence structure in bed load transport and bed form mechanics. Water Resources Research 31: 2071-2086. https://doi.org/10.1029/ 95WR00976
Nepl HM. 1999. Drag, turbulence, and diffusion in flow through emergent vegetation. Water Resources Research 35: 479-489.
Nielsen P. 1984. On the motion of suspended sand particles. Journal of Geophysical Research: Oceans 89: 616-626.
Nielsen P. 1993. Turbulence effects on the settling of suspended particles. SEPM Journal of Sedimentary Research 63(5), 835-838. https:// doi.org/10.1306/D4267C1C-2B26-11D7-8648000102C1865D
Nittrouer JA, Allison MA, Campanella R. 2008. Bedform transport rates for the lowermost Mississippi River. Journal of Geophysical Research 113(F03004), 1-16. https://doi.org/10.1029/2007JF000795
Omidyeganeh M, Piomelli U. 2013. Large-eddy simulation of three-dimensional dunes in a steady, unidirectional flow. Part 2. Flow structures. Journal of Fluid Mechanics 734: 509-534.
Ortiz AC, Ashton A, Nepf H. 2013. Mean and turbulent velocity fields near rigid and flexible plants and the implications for deposition. Journal of Geophysical Research: Earth Surface 118: 2585-2599.
Paarlberg AJ, Dohmen-Janssen CM, Hulscher SJ, Termes P. 2009. Modeling river dune evolution using a parameterization of flow separation. Journal of Geophysical Research: Earth Surface 114(F01014), 1-17. https://doi.org/10.1029/2007JF000910
Paola C, Twilley RR, Edmonds DA, Kim W, Mohrig D, Parker G, Viparelli E, Voller VR. 2011. Natural processes in delta restoration: application to the Mississippi Delta. Annual Review of Marine Science 3: 67-91. https://doi.org/10.1146/annurev-marine-120709142856
Papanicolaou AN, Diplas P, Evaggelopoulos N, Fotopoulos S. 2002. Stochastic incipient motion criterion for spheres under various bed packing conditions. Journal of Hydraulic Engineering 128: 369-380.
Parsons DR, Best J. 2013. Bedforms: views and new perspectives from the Third International Workshop on Marine and River Dune Dynamics (MARID3). Earth Surface Processes and Landforms 38: 319-329.
Parsons DR, Best JL, Orfeo O, Hardy RJ, Kostaschuk R, Lane SN. 2005. Morphology and flow fields of three-dimensional dunes, Rio Paraná, Argentina: results from simultaneous multibeam echo sounding and acoustic Doppler current profiling. Journal of Geophysical Research: Earth Surface 110(F04503), 1-9. https://doi.org/10.1029/ 2004JF000231
Pasiok R, Stilger-Szydło E. 2010. Sediment particles and turbulent flow simulation around bridge piers. Archives of Civil and Mechanical Engineering 10: 67-79.
Prent MT, Hickin EJ. 2001. Annual regime of bedforms, roughness and flow resistance, Lillooet River, British Columbia, BC. Geomorphology 41: 369-390.
Ramirez MT, Allison MA. 2013. Suspension of bed material over sand bars in the Lower Mississippi River and its implications for Mississippi delta environmental restoration. Journal of Geophysical Research: Earth Surface 110: 1085-1104. https://doi.org/10.1002/jgrl. 20075
Raudkivi AJ. 1998. Loose Boundary Hydraulics. A.A. Balkema: Rotterdam.
Reesink AJH, Bridge JS. 2007. Influence of superimposed bedforms and flow unsteadiness on formation of cross strata in dunes and unit bars. Sedimentary Geology 202: 281-296.
van Rijn LC. 1984. Sediment transport, part II: suspended load transport. Journal of Hydraulic Engineering 110: 1613-1641.
van Rijn LC. 2007. Unified view of sediment transport by currents and waves. II: Suspended Transport. Journal of Hydraulic Engineering 133: 668-689. https://doi.org/10.1061/(ASCE)0733$9429(2007) 133: 6(668)vanRijnLC,TanGL.1985.SUTRENCH−model:two−dimensionalverticalmathematicalmodelforsedimentationindredgedchannelsandtrenchesbycurrentsandwaves.RijkswaterstaatCommunications:TheHague.RoodKM,HickinEJ.1989.Suspended−sedimentconcentrationandcalibreinrelationtosurface−flowstructureinSquamishRiverestuary,southwesternBritishColumbia.CanadianJournalofEarthSciences26:2172−2176.https://doi.org/10.1139/e89−183SchmeeckleMW.2014.Numericalsimulationofturbulenceandsedimenttransportofmediumsand.JournalofGeophysicalResearch:EarthSurface119:1240−1262.SchmeeckleMW.2015.Theroleofvelocity,pressure,andbedstressfluctuationsinbedloadtransportoverbedforms:numericalsimulationdownstreamofabackward−lacingstep.EarthSurfaceDynamics3:105−112.https://doi.org/10.5194/esurf−3−105−2015SchmeeckleMW,ShimizuY,HoshiK,BabaH,IkezakiS.1999.Turbulentstructuresandsuspendedsedimentovertwo−dimensionaldunes.JournalofHydraulicEngineering134:261−270.ShamsM,AhmadiG,SmithDH.2002.Computationalmodelingofflowandsedimenttransportanddepositioninmeanderingrivers.AdvancesinWaterResources25:689−699.ShengYP,VillaretC.1989.Modelingtheeffectofsuspendedsedimentstratificationonbottomexchangeprocesses.JournalofGeophysicalResearch:Oceans94:14429−14444.ShugarDH,KostaschukR,BestJL,ParsonsDR,LaneSN,OrfeoO,HardyRJ.2010.Ontherelationshipbetweenflowandsuspendedsedimenttransportoverthecrestofasanddune,RioParanaˊ,Argentina.Sedimentology57:252−272.SimonsDB,RichardsonEV.1961.Formsofbedroughnessinalluvialchannel.JournaloftheHydraulicsDivision87:87−105.SimonsDB,RichardsonEV,NordinJrCF.1965.BedloadEquationforRipplesandDunes.U.S.GeologicalSurveyProfessionalPaper:Washington,D.C.SmithJD,McLeanSR.1977.Spatiallyaveragedflowoverawavysurface.JournalofGeophysicalResearch82:1735−1746.SpalartP,AllmarasS.1994.Aone−equationturbulencemodelforaerodynamicflows.LaRechercheAerospatiale1:5−21.StommelH.1949.Trajectoriesofsmallbodiessinkingslowlythroughconvectioncells.JournalofMarineResearch8:24−29.StoutJE,AryaSP,GenikhovichEL.1995.Theeffectofnonlineardragonthemotionandsettlingvelocityofheavyparticles.JournaloftheAtmosphericSciences52:3836−3848.https://doi.org/10.1175/1520−0469(1995)052<3836:TEONDO>2.0 . \mathrm{CO}$; 2
Tominaga Y, Stathopoulos T. 2011. CFD modeling of pollution dispersion in a street canyon: comparison between LES and RANS. Journal of Wind Engineering and Industrial Aerodynamics 99: 340-348.
Tooby PF, Wick GL, Isaacs JD. 1977. The motion of a small sphere in a rotating velocity field: a possible mechanism for suspending particles in turbulence. Journal of Geophysical Research 82: 2096-2100.
Umlauf L, Burchard H. 2003. A generic length-scale equation for geophysical turbulence models. Journal of Marine Research 61: 235-265. https://doi.org/10.1357/002224003322005087
VanSickle J, Beschta RL. 1983. Supply-based models of suspended sediment transport in streams. Water Resources Research 19: 768−778.
Vargas-Luna A, Crosato A, Uijttewaal WS. 2015. Effects of vegetation on flow and sediment transport: comparative analyses and validation of predicting models. Earth Surface Processes and Landforms 40: 157−176.
Venditti JG. 2007. Turbulent flow and drag over fixed two- and three-dimensional dunes. Journal of Geophysical Research 112 (F04008), 1-21. https://doi.org/10.1029/2006JF000650
Venditti JG, Bauer BO. 2005. Turbulent flow over a dune: Green River, Colorado. Earth Surface Processes and Landforms: The Journal of the British Geomorphological Research Group 30: 289-304.
Venditti JG, Bennett SJ. 2000. Spectral analysis of turbulent flow and suspended sediment transport over fixed dunes. Journal of Geophysical Research: Oceans 105: 22035-22047. https://doi.org/10.1029/ 2000JC900094
Wang Y. 2013. Development of a numerical tool to predict hydrodynamics, temperature and TDG in hydropower flows. PhD thesis, University of lowa, Ames, IA.
Wang L-P, Maxey MR. 1993. Settling velocity and concentration distribution of heavy particles in homogeneous isotropic turbulence. Journal of Fluid Mechanics 256: 27-68. https://doi.org/10.1017/ S0022112093002708
Wang Y, Politano M, Laughery R, Weber L. 2015. Model development in OpenFOAM to predict spillway jet regimes. Journal of Applied Water Engineering and Research 3: 80-94.
Wei T, Willmarth WW. 1991. Examination of v-velocity fluctuations in a turbulent channel flow in the context of sediment transport. Journal of Fluid Mechanics 223: 241-252.
Wren DG, Kuhnle RA, Wilson CG. 2007. Measurements of the relationship between turbulence and sediment in suspension over mobile sand dunes in a laboratory flume. Journal of Geophysical Research: Earth Surface 112(F03009), 1-14.
Yager EM, Schmeeckle MW. 2013. The influence of vegetation on turbulence and bed load transport. Journal of Geophysical Research: Earth Surface 118: 1585-1601.
Yang S-Q. 2007. Turbulent transfer mechanism in sediment-laden flow. Journal of Geophysical Research 112(F01005), 1-14. https://doi.org/ 10.1029/2005JF000452
Yang JQ, Nepf HM. 2018. A turbulence-based bed-load transport model for bare and vegetated channels. Geophysical Research Letters 45: 10428-10436.
Zedler EA, Street RL. 2001. Large-eddy simulation of sediment transport: currents over ripples. Journal of Hydraulic Engineering 127: 444-452. https://doi.org/10.1061/(ASCE)0733-9429(2001)127:6(444)
Zhao T, Dai F, Xu N. 2017. Coupled DEM-CFD investigation on the formation of landslide dams in narrow rivers. Landslides 14: 189-201. 1
Supporting Information
Additional supporting information may be found online in the Supporting Information section at the end of the article.
Supporting Information S1
References (133)
Allen JRL. 1978. Polymodal dune assemblages: an interpretation in terms of dune creation-destruction in periodic flows. Sedimentary Geology 20: 17-28.
Allison MA, Yuill BT, Meselhe EA, Marsh JK, Kolker AS, Ameen AD. 2017. Observational and numerical particle tracking to examine sed- iment dynamics in a Mississippi River delta diversion. Estuarine, Coastal and Shelf Science 194: 97-108. https://doi.org/10.1016/j. ecss.2017.06.004
Alvarez LV, Schmeeckle MW, Grams PE. 2017. A detached eddy simu- lation model for the study of lateral separation zones along a large canyon-bound river. Journal of Geophysical Research: Earth Surface 122: 25-49.
Amsler ML, Schreider MI. 1999. Dune height prediction at floods in the Paraná River, Argentina. In River Sedimentation: Theory and Applica- tions, Jayewardena AW, Lee JHW, Wang ZY (eds). A.A. Balkema: Rotterdam; 615-620.
Atkins R, Soulsby RL, Waters CB, Oliver N. 1989. Field measurements of sediment suspension above bedforms in a sandy estuary. Techni- cal Report. Hydraulics Research: Wallingford.
Badano ND, Sabarots Gerbec M, Re M, Menéndez AN. 2012. A coupled hydrosedimentologic model to assess the advance of the Parana River Delta Front. In Proceedings of the Sixth International Conference on Fluvial Hydraulics, San Jose, Costa Rica.
B. YUILL ET AL.
Bagnold RA. 1966. An Approach to the Sediment Transport Problem from General Physics. US Government Printing Office: Washington, D.C. Bennett SJ, Best JL. 1995. Mean flow and turbulence structure over fixed, two-dimensional dunes: implications for sediment transport and bedform stability. Sedimentology 42: 491-513. https://doi.org/ 10.1111/j.1365-3091.1995.tb00386.x
Bennett SJ, Bridge JS, Best JL. 1998. Fluid and sediment dynamics of upper stage plane beds. Journal of Geophysical Research: Oceans 103: 1239-1274. https://doi.org/10.1029/97JC02764
Best J. 2005. The fluid dynamics of river dunes: a review and some future research directions. Journal of Geophysical Research: Earth Surface 110: 1-21. https://doi.org/10.1029/2004JF000218
Best J, Kostaschuk R. 2002. An experimental study of turbulent flow over a low-angle dune. Journal of Geophysical Research: Oceans 107: 18-1-18-19.
Bhaganagar K, Hsu T-J. 2009. Direct numerical simulations of flow over two-dimensional and three-dimensional ripples and implication to sediment transport: steady flow. Coastal Engineering 56: 320-331. https://doi.org/10.1016/j.coastaleng.2008.09.010
Bijker EW, Van Hijum E, Vellinga P. 1976. Sand transport by waves. In Coastal Engineering Proceedings.
Bradley RW, Venditti JG. 2019. The growth of dunes in rivers. Journal of Geophysical Research: Earth Surface 124: 548-566.
Bradley RW, Venditti JG, Kostaschuk RA, Church M, Hendershot M, Allison MA. 2013. Flow and sediment suspension events over low-angle dunes: Fraser Estuary, Canada. Journal of Geophysical Research: Earth Surface 118: 1693-1709.
Bridge JS, Best JL. 1988. Flow, sediment transport and bedform dynam- ics over the transition from dunes to upper-stage plane beds: implica- tions for the formation of planar laminae. Sedimentology 35: 753-763.
Brucato A, Grisafi F, Montante G. 1998. Particle drag coefficients in tur- bulent fluids. Chemical Engineering Science 53: 3295-3314.
Carling PA, Williams JJ, Golz E, Kelsey AD. 2000. The morphodynamics of fluvial sand dunes in the River Rhine, near Mainz, Germany. II. Hydrodynamics and sediment transport. Sedi- mentology 47(1), 253-278.
Cellino M, Lemmin U. 2004. Influence of coherent flow structures on the dynamics of suspended sediment transport in open-channel flow. Journal of Hydraulic Engineering 130: 1077-1088.
Chakrabarti A, Chen Q, Smith HD, Liu D. 2016. Large eddy simulation of unidirectional and wave flows through vegetation. Journal of Engi- neering Mechanics 142: 1-8, 04016048.
Chang YS. 2004. Suspended sediment and hydrodynamics above mildly sloped long wave ripples. Journal of Geophysical Research 109: 1-16. https://doi.org/10.1029/2003JC001900
Chang YS, Park Y-G. 2016. Suspension of sediment particles over a rip- ple due to turbulent convection under unsteady flow conditions. Ocean Science Journal 51: 127-135. https://doi.org/10.1007/ s12601-016-0011-2
Chang Y, Scotti A. 2003. Entrainment and suspension of sediments into a turbulent flow over ripples. Journal of Turbulence 4: 1-22. https:// doi.org/10.1088/1468-5248/4/1/019
Chang WY, Constantinescu G, Tsai WF, Lien HC. 2011. Coherent struc- ture dynamics and sediment erosion mechanisms around an in-stream rectangular cylinder at low and moderate angles of attack. Water Resources Research 47(W12532), 1-16. https://doi.org/ 10.1029/2011WR010586.
Chen X, Cardenas MB, Chen L. 2015. Three-dimensional versus two-dimensional bed form-induced hyporheic exchange. Water Resources Research 51: 2923-2936.
Davies AG, Thorne PD. 2016. On the suspension of graded sediment by waves above ripples: inferences of convective and diffusive pro- cesses. Continental Shelf Research 112: 46-67. https://doi.org/ 10.1016/j.csr.2015.10.006
Dietrich WE. 1982. Settling velocity of natural particles. Water Resources Research 18: 1615-1626. https://doi.org/10.1029/ WR018i006p01615
Dietrich WE, Smith JD. 1984. Bed load transport in a river meander. Water Resources Research 20: 1355-1380.
Duan JG, Julien PY. 2010. Numerical simulation of meandering evolu- tion. Journal of Hydrology 391: 34-46.
Dubief Y, Delcayre F. 2000. On coherent-vortex identification in turbu- lence. Journal of Turbulence 1: 1-22. https://doi.org/10.1088/1468- 5248/1/1/011
Engelund F, Fredsøe J. 1976. A sediment transport model for straight alluvial channels. Hydrology Research 7: 293-306.
Fornari W, Picano F, Brandt L. 2016. Sedimentation of finite-size spheres in quiescent and turbulent environments. Journal of Fluid Mechanics 788: 640-669. https://doi.org/10.1017/jfm.2015.698
Fung JCH. 1998. Effect of nonlinear drag on the settling velocity of par- ticles in homogeneous isotropic turbulence. Journal of Geophysical Research: Oceans 103: 27905-27917. https://doi.org/10.1029/ 98JC02822
Gabel SL. 1993. Geometry and kinematics of dunes during steady and unsteady flows in the Calamus River, Nebraska, USA. Sedimentology 40: 237-269.
Garcia M, Parker G. 1991. Entrainment of bed sediment into suspen- sion. Journal of Hydraulic Engineering 117: 414-435.
Gelfenbaum G, Smith JD. 1986. Experimental evaluation of a general- ized suspended-sediment transport theory. In Shelf Sands and Sand- stones, Knight RJ, McLean JR (eds). Canadian Society of Petroleum Geologists: Calgary; 133-134.
Grass AJ. 1971. Structural features of turbulent flow over smooth and rough boundaries. Journal of Fluid Mechanics 50: 233-255.
Grigoriadis DGE, Balaras E, Dimas AA. 2009. Large-eddy simulations of unidirectional water flow over dunes. Journal of Geophysical Research 114(F02022), 1-19. https://doi.org/10.1029/2008JF001014
Hand BM, Bartberger CE. 1988. Leeside sediment fallout patterns and the stability of angular bedforms. Journal of Sedimentary Research 58: 33-43. https://doi.org/10.1306/212F8D05-2B24-11D7- 8648000102C1865D
Hendershot ML, Venditti JG, Bradley RW, Kostaschuk RA, Church M, Allison MA. 2016. Response of low-angle dunes to variable flow. Sedimentology 63: 743-760.
Herbert CM, Alexander J. 2018. Bottomset architecture formed in the troughs of dunes and unit bars. Journal of Sedimentary Research 88: 522-553.
Howard AD. 1971. Simulation of stream networks by headword growth and branching. Geographical Analysis 3: 29-50.
Hurther D, Lemmin U. 2003. Turbulent particle flux and momentum flux statistics in suspension flow. Water Resources Research 39(5), 1139-1150. https://doi.org/10.1029/2001WR001113.
Jerolmack DJ, Mohrig D. 2005. A unified model for subaqueous bed form dynamics. Water Resources Research 41(W12421), 1-10. https://doi.org/10.1029/2005WR004329.
Johannesson H, Parker G. 1989. Linear theory of river meanders. River Meandering 12: 181-213.
Johns B, Xing J. 1993. Three-dimensional modelling of the free surface turbulent flow of water over a bedform. Continental Shelf Research 13: 705-721.
Keylock CJ, Constantinescu G, Hardy RJ. 2012. The application of com- putational fluid dynamics to natural river channels: eddy resolving versus mean flow approaches. Geomorphology 179: 1-20. https:// doi.org/10.1016/j.geomorph.2012.09.006
Keylock CJ, Lane SN, Richards KS. 2014. Quadrant/octant sequencing and the role of coherent structures in bed load sediment entrainment. Journal of Geophysical Research: Earth Surface 119: 264-286. https:// doi.org/10.1002/2012JF002698
Khosronejad A, Sotiropoulos F. 2014. Numerical simulation of sand waves in a turbulent open channel flow. Journal of Fluid Mechanics 753: 150-216. https://doi.org/10.1017/jfm.2014.335
Kirchner JW, Dietrich WE, Iseya F, Ikeda H. 1990. The variability of crit- ical shear stress, friction angle, and grain protrusion in water-worked sediments. Sedimentology 37: 647-672.
Kostaschuk R. 2000. A field study of turbulence and sediment dynamics over subaqueous dunes with flow separation. Sedimentology 47: 519-531.
Kostaschuk R, Best J. 2005. Response of sand dunes to variations in tidal flow: Fraser Estuary. Canada. Journal of Geophysical Research: Earth Surface 110(F04S04), 1-10. https://doi.org/10.1029/ 2004JF000176
Kostaschuk RA, Church MA. 1993. Macroturbulence generated by dunes: Fraser River, Canada. Sedimentary Geology 85: 25-37. https://doi.org/10.1016/0037-0738(93)90073-E SAND SETTLING THROUGH BEDFORM-GENERATED TURBULENCE IN RIVERS Kostaschuk R, Villard P. 1996. Flow and sediment transport over large subaqueous dunes: Fraser River, Canada. Sedimentology 43: 849-863.
Kostaschuk RA, Villard PV. 1999. Turbulent sand suspension over dunes. In Fluvial Sedimentology VI, Vol. 28; 3-14.
Kostaschuk R, Shugar D, Best J, Parsons D, Lane S, Hardy R, Orfeo O. 2009. Suspended sediment transport and deposition over a dune: Río Paraná, Argentina. Earth Surface Processes and Landforms 34: 1605-1611.
Kreplin H-P, Eckelmann H. 1979. Behavior of the three fluctuating velocity components in the wall region of a turbulent channel flow. The Physics of Fluids 22: 1233-1239.
Kwoll E, Winter C, Becker M. 2013. Intermittent suspension and trans- port of fine sediment over natural tidal bedforms. In Coherent Flow Structures at Earth's Surface, Venditti JG, Best JL, Church M, Hardy RJ (eds). Wiley-Blackwell: Oxford; 231-242.
Kwoll E, Venditti JG, Bradley RW, Winter C. 2016. Flow structure and resistance over subaqueous high-and low-angle dunes. Journal of Geophysical Research: Earth Surface 121: 545-564. https://doi.org/ 10.1002/2015JF003637
Lai YG, Bandrowski DJ. 2014. Large wood flow hydraulics: a 3D modelling approach. In Proceedings of Bold Visions for Environmen- tal Modelling, San Diego, CA.
Lapointe M. 1992. Burst-like sediment suspension events in a sand bed river. Earth Surface Processes and Landforms 17: 253-270. https:// doi.org/10.1002/esp.3290170305
Leeder MR. 1983. On the dynamics of sediment suspension by residual Reynolds stresses-confirmation of Bagnold's theory. Sedimentology 30: 485-491.
Leeder MR, Gray TE, Alexander J. 2005. Sediment suspension dynam- ics and a new criterion for the maintenance of turbulent suspensions. Sedimentology 52: 683-691.
Lefebvre A. 2019. Three-dimensional flow above river bedforms: insights from numerical modeling of a natural dune field (Río Paraná, Argentina). Journal of Geophysical Research: Earth Surface 124: 2241-2264.
Lefebvre A, Winter C. 2016. Predicting bed form roughness: the influ- ence of lee side angle. Geo-Marine Letters 36: 121-133.
Lefebvre A, Paarlberg AJ, Winter C. 2014. Flow separation and shear stress over angle-of-repose bed forms: a numerical investigation. Water Resources Research 50: 986-1005. https://doi.org/10.1002/ 2013WR014587
Lelouvetel J, Bigillon F, Doppler D, Vinkovic I, Champagne J-Y. 2009. Experimental investigation of ejections and sweeps involved in parti- cle suspension. Water Resources Research 45(W02416), 1-15. https://doi.org/10.1029/2007WR006520
Luchik TS, Tiederman WG. 1987. Timescale and structure of ejections and bursts in turbulent channel flows. Journal of Fluid Mechanics 174: 529-552.
Marchioli C, Armenio V, Salvetti MV, Soldati A. 2006. Mechanisms for deposition and resuspension of heavy particles in turbulent flow over wavy interfaces. Physics of Fluids 18: 1-16, 025102. https://doi.org/ 10.1063/1.2166453
McElroy B, Mohrig D. 2009. Nature of deformation of sandy bed forms. Journal of Geophysical Research 114(F00A04), 1-13. https://doi.org/ 10.1029/2008JF001220
McLean SR. 1992. On the calculation of suspended load for noncohesive sediments. Journal of Geophysical Research: Oceans 97: 5759-5770.
McLean SR, Nelson JM, Wolfe SR. 1994. Turbulence structure over two-dimensional bed forms: implications for sediment transport. Journal of Geophysical Research 99(C6), 12729-12747. https://doi. org/10.1029/94JC00571
McQuivey RS. 1973. Summary of Turbulence Data from Rivers, Con- veyance Channels, and Laboratory Flumes: Turbulence in Water. US Government Printing Office: Washington, D.C.
Middleton GV, Southard JB. 1984. Mechanics of sediment movement. Short Course No. 3. Society of Economic Paleontologists and Miner- alogists: Providence, RI.
Mohrig D, Smith JD. 1996. Predicting the migration rates of subaque- ous dunes. Water Resources Research 32: 3207-3217. https://doi. org/10.1029/96WR01129
Molinas A, Wu B. 2001. Transport of sediment in large sand-bed rivers. Journal of Hydraulic Research 39: 135-146.
Müller A, Gyr A. 1986. On the vortex formation in the mixing layer behind dunes. Journal of Hydraulic Research 24: 359-375.
Murray SP. 1970. Settling velocities and vertical diffusion of particles in turbulent water. Journal of Geophysical Research 75: 1647-1654. https://doi.org/10.1029/JC075i009p01647
Murray AB, Paola C. 1994. A cellular model of braided rivers. Nature 371: 54-57.
Nagata N, Hosoda T, Muramoto Y. 2000. Numerical analysis of river channel processes with bank erosion. Journal of Hydraulic Engineer- ing 126: 243-252.
Naqshband S, Ribberink JS, Hurther D, Hulscher SJ. 2014. Bed load and suspended load contributions to migrating sand dunes in equilib- rium. Journal of Geophysical Research: Earth Surface 119: 1043-1063.
Nelson JM, McLean SR, Wolfe SR. 1993. Mean flow and turbulence fields over two-dimensional bed forms. Water Resources Research 29: 3935-3953. https://doi.org/10.1029/93WR01932
Nelson JM, Shreve RL, McLean SR, Drake TG. 1995. Role of near-bed turbulence structure in bed load transport and bed form mechanics. Water Resources Research 31: 2071-2086. https://doi.org/10.1029/ 95WR00976
Nepf HM. 1999. Drag, turbulence, and diffusion in flow through emer- gent vegetation. Water Resources Research 35: 479-489.
Nielsen P. 1984. On the motion of suspended sand particles. Journal of Geophysical Research: Oceans 89: 616-626.
Nielsen P. 1993. Turbulence effects on the settling of suspended parti- cles. SEPM Journal of Sedimentary Research 63(5), 835-838. https:// doi.org/10.1306/D4267C1C-2B26-11D7-8648000102C1865D
Nittrouer JA, Allison MA, Campanella R. 2008. Bedform transport rates for the lowermost Mississippi River. Journal of Geophysical Research 113(F03004), 1-16. https://doi.org/10.1029/2007JF000795
Omidyeganeh M, Piomelli U. 2013. Large-eddy simulation of three-dimensional dunes in a steady, unidirectional flow. Part 2. Flow structures. Journal of Fluid Mechanics 734: 509-534.
Ortiz AC, Ashton A, Nepf H. 2013. Mean and turbulent velocity fields near rigid and flexible plants and the implications for deposition. Journal of Geophysical Research: Earth Surface 118: 2585-2599.
Paarlberg AJ, Dohmen-Janssen CM, Hulscher SJ, Termes P. 2009. Modeling river dune evolution using a parameterization of flow sep- aration. Journal of Geophysical Research: Earth Surface 114(F01014), 1-17. https://doi.org/10.1029/2007JF000910
Paola C, Twilley RR, Edmonds DA, Kim W, Mohrig D, Parker G, Viparelli E, Voller VR. 2011. Natural processes in delta restoration: application to the Mississippi Delta. Annual Review of Marine Sci- ence 3: 67-91. https://doi.org/10.1146/annurev-marine-120709- 142856
Papanicolaou AN, Diplas P, Evaggelopoulos N, Fotopoulos S. 2002. Stochastic incipient motion criterion for spheres under various bed packing conditions. Journal of Hydraulic Engineering 128: 369-380.
Parsons DR, Best J. 2013. Bedforms: views and new perspectives from the Third International Workshop on Marine and River Dune Dynam- ics (MARID3). Earth Surface Processes and Landforms 38: 319-329.
Parsons DR, Best JL, Orfeo O, Hardy RJ, Kostaschuk R, Lane SN. 2005. Morphology and flow fields of three-dimensional dunes, Rio Paraná, Argentina: results from simultaneous multibeam echo sounding and acoustic Doppler current profiling. Journal of Geophysical Research: Earth Surface 110(F04S03), 1-9. https://doi.org/10.1029/ 2004JF000231
Pasiok R, Stilger-Szydło E. 2010. Sediment particles and turbulent flow simulation around bridge piers. Archives of Civil and Mechanical Engineering 10: 67-79.
Prent MT, Hickin EJ. 2001. Annual regime of bedforms, roughness and flow resistance, Lillooet River, British Columbia, BC. Geomorphol- ogy 41: 369-390.
Ramirez MT, Allison MA. 2013. Suspension of bed material over sand bars in the Lower Mississippi River and its implications for Mississippi delta environmental restoration. Journal of Geophysical Research: Earth Surface 118: 1085-1104. https://doi.org/10.1002/jgrf.20075
Raudkivi AJ. 1998. Loose Boundary Hydraulics. A.A. Balkema: Rotterdam.
B. YUILL ET AL.
Reesink AJH, Bridge JS. 2007. Influence of superimposed bedforms and flow unsteadiness on formation of cross strata in dunes and unit bars. Sedimentary Geology 202: 281-296.
van Rijn LC. 1984. Sediment transport, part II: suspended load trans- port. Journal of Hydraulic Engineering 110: 1613-1641.
van Rijn LC. 2007. Unified view of sediment transport by currents and waves. II: Suspended Transport. Journal of Hydraulic Engineering 133: 668-689. https://doi.org/10.1061/(ASCE)0733- 9429(2007)133:6(668)
van Rijn LC, Tan GL. 1985. SUTRENCH-model: two-dimensional verti- cal mathematical model for sedimentation in dredged channels and trenches by currents and waves. Rijkswaterstaat Communications: The Hague.
Rood KM, Hickin EJ. 1989. Suspended-sediment concentration and calibre in relation to surface-flow structure in Squamish River estu- ary, southwestern British Columbia. Canadian Journal of Earth Sci- ences 26: 2172-2176. https://doi.org/10.1139/e89-183
Schmeeckle MW. 2014. Numerical simulation of turbulence and sedi- ment transport of medium sand. Journal of Geophysical Research: Earth Surface 119: 1240-1262.
Schmeeckle MW. 2015. The role of velocity, pressure, and bed stress fluctuations in bed load transport over bed forms: numerical simula- tion downstream of a backward-facing step. Earth Surface Dynamics 3: 105-112. https://doi.org/10.5194/esurf-3-105-2015
Schmeeckle MW, Shimizu Y, Hoshi K, Baba H, Ikezaki S. 1999. Turbu- lent structures and suspended sediment over two-dimensional dunes. Journal of Hydraulic Engineering 134: 261-270.
Shams M, Ahmadi G, Smith DH. 2002. Computational modeling of flow and sediment transport and deposition in meandering rivers. Advances in Water Resources 25: 689-699.
Sheng YP, Villaret C. 1989. Modeling the effect of suspended sediment stratification on bottom exchange processes. Journal of Geophysical Research: Oceans 94: 14429-14444.
Shugar DH, Kostaschuk R, Best JL, Parsons DR, Lane SN, Orfeo O, Hardy RJ. 2010. On the relationship between flow and suspended sediment transport over the crest of a sand dune, Río Paraná, Argentina. Sedimentology 57: 252-272.
Simons DB, Richardson EV. 1961. Forms of bed roughness in alluvial channel. Journal of the Hydraulics Division 87: 87-105.
Simons DB, Richardson EV, Nordin Jr CF. 1965. Bedload Equation for Ripples and Dunes. U.S. Geological Survey Professional Paper: Washington, D.C.
Smith JD, McLean SR. 1977. Spatially averaged flow over a wavy sur- face. Journal of Geophysical Research 82: 1735-1746.
Spalart P, Allmaras S. 1994. A one-equation turbulence model for aero- dynamic flows. La Recherche Aerospatiale 1: 5-21.
Stommel H. 1949. Trajectories of small bodies sinking slowly through convection cells. Journal of Marine Research 8: 24-29.
Stout JE, Arya SP, Genikhovich EL. 1995. The effect of nonlinear drag on the motion and settling velocity of heavy particles. Journal of the Atmospheric Sciences 52: 3836-3848. https://doi.org/10.1175/ 1520-0469(1995)052<3836:TEONDO>2.0.CO;
Tominaga Y, Stathopoulos T. 2011. CFD modeling of pollution disper- sion in a street canyon: comparison between LES and RANS. Journal of Wind Engineering and Industrial Aerodynamics 99: 340-348.
Tooby PF, Wick GL, Isaacs JD. 1977. The motion of a small sphere in a rotating velocity field: a possible mechanism for suspending particles in turbulence. Journal of Geophysical Research 82: 2096-2100.
Umlauf L, Burchard H. 2003. A generic length-scale equation for geo- physical turbulence models. Journal of Marine Research 61: 235-265. https://doi.org/10.1357/002224003322005087
VanSickle J, Beschta RL. 1983. Supply-based models of suspended sediment transport in streams. Water Resources Research 19: 768-778.
Vargas-Luna A, Crosato A, Uijttewaal WS. 2015. Effects of vegetation on flow and sediment transport: comparative analyses and validation of predicting models. Earth Surface Processes and Landforms 40: 157-176.
Venditti JG. 2007. Turbulent flow and drag over fixed two-and three-dimensional dunes. Journal of Geophysical Research 112 (F04008), 1-21. https://doi.org/10.1029/2006JF000650
Venditti JG, Bauer BO. 2005. Turbulent flow over a dune: Green River, Colorado. Earth Surface Processes and Landforms: The Journal of the British Geomorphological Research Group 30: 289-304.
Venditti JG, Bennett SJ. 2000. Spectral analysis of turbulent flow and suspended sediment transport over fixed dunes. Journal of Geophys- ical Research: Oceans 105: 22035-22047. https://doi.org/10.1029/ 2000JC900094
Wang Y. 2013. Development of a numerical tool to predict hydrody- namics, temperature and TDG in hydropower flows. PhD thesis, Uni- versity of Iowa, Ames, IA.
Wang L-P, Maxey MR. 1993. Settling velocity and concentration distri- bution of heavy particles in homogeneous isotropic turbulence. Jour- nal of Fluid Mechanics 256: 27-68. https://doi.org/10.1017/ S0022112093002708
Wang Y, Politano M, Laughery R, Weber L. 2015. Model development in OpenFOAM to predict spillway jet regimes. Journal of Applied Water Engineering and Research 3: 80-94.
Wei T, Willmarth WW. 1991. Examination of v-velocity fluctuations in a turbulent channel flow in the context of sediment transport. Journal of Fluid Mechanics 223: 241-252.
Wren DG, Kuhnle RA, Wilson CG. 2007. Measurements of the relation- ship between turbulence and sediment in suspension over mobile sand dunes in a laboratory flume. Journal of Geophysical Research: Earth Surface 112(F03009), 1-14.
Yager EM, Schmeeckle MW. 2013. The influence of vegetation on tur- bulence and bed load transport. Journal of Geophysical Research: Earth Surface 118: 1585-1601.
Yang S-Q. 2007. Turbulent transfer mechanism in sediment-laden flow. Journal of Geophysical Research 112(F01005), 1-14. https://doi.org/ 10.1029/2005JF000452
Yang JQ, Nepf HM. 2018. A turbulence-based bed-load transport model for bare and vegetated channels. Geophysical Research Letters 45: 10428-10436.
Zedler EA, Street RL. 2001. Large-eddy simulation of sediment transport: currents over ripples. Journal of Hydraulic Engineering 127: 444-452. https://doi.org/10.1061/(ASCE)0733-9429(2001)127:6(444)
Zhao T, Dai F, Xu N. 2017. Coupled DEM-CFD investigation on the for- mation of landslide dams in narrow rivers. Landslides 14: 189-201. 1
Turbulence measurements from the flow over two-dimensional fixed dune shapes are presented along with analysis and discussion of the ramifications of the observations for transport of sediment as bed load over bed forms. The spatial structure of the local transport rate determines the shape and stability of bed forms such as tipples and dunes, and the transport of sediment is a highly nonlinear process that is profoundly affected by the statistics of the temporal fluctuations in the near-bed flow field. The measurements presented herein show strong spatial evolution of the joint probability distribution of the streamwise and bednormal fluctuating velocity components. Unlike measurements in uniform boundary layers, these distributions and the higher moments of the velocity do not scale with the local shear velocity,indicating that it is probably inappropriate to use the shear stress to characterize the sediment flux. This conclusion is supported by observations of sediment flux over a dune. Introduction Ripples and dunes play a critical role in the complex interrelation between flow, sediment flux, and bed morphology in many natural flows over erodible beds. In spite of the importance of these features, our ability to predict their occurrence, their sizes and shapes, and their effect on the flow is limited. This is primarily a result of our incomplete understanding of (1) the flow adjustments caused by bed forms, (2) the response of sediment to local fluid forcing, and (3) the subsequent evolution of the bed forms. These processes are interrelated, and none can be investigated completely independently of the others: the local rate of sediment transport is dependent on the near-bed flow field, which is controlled by the shape of the bottom topography, which in turn is altered if there is a local convergence or divergence of sediment flux. In addition, sediment motion acts to modify the flow field by enhancing roughness in the case of bed load and by damping turbulence in the case of suspended-sediment load. Furthermore, the shape of the bottom can alter the sediment load characteristics by inducing gravitational effects associated with bed slope. This complex, highly nonlinear interdependence often leads to responses of the system that are
The stability of bedforms in mixtures of gravel and sand is not well understood. Two bedform types are characteristic: flowparallel sand ribbons and flow-transverse barchans. Flume experiments and field data presented here show that gradual transitions exist from sand ribbons to barchans, and from barchans to fully developed dunes. Barchans and sand ribbons occur when not enough transportable sediment is available for the formation of fully developed ripples or dunes. The reason is that a part of the bed sediment is immobile, e.g., with an armor layer, which limits the sediment supply and thus the volume of sediment available for the formation of bedforms.
An experimental field study of flows over bedforms was conducted on the Missouri River near St. Charles, Missouri. Detailed velocity data were collected under two different flow conditions along bedforms in this sand-bed river. The large river-scale data reflect flow characteristics similar to those of laboratory-scale flows, with flow separation occurring downstream of the bedform crest and flow reattachment on the stoss side of the next downstream bedform. Wave-like responses of the flow to the bedforms were detected, with the velocity decreasing throughout the flow depth over bedform troughs, and the velocity increasing over bedform crests. Local and spatially averaged velocity distributions were logarithmic for both datasets. The reach-wise spatially averaged vertical-velocity profile from the standard velocity-defect model was evaluated. The vertically averaged mean flow velocities for the velocity-defect model were within 5% of the measured values and estimated spatially averaged point velocities were within 10% for the upper 90% of the flow depth. The velocity-defect model, neglecting the wake function, was evaluated and found to estimate the vertically averaged mean velocity within 1% of the measured values. RÉSUMÉ Une étude expérimentale en nature des écoulements sur des formes de lit a été conduite sur le fleuve Missouri près de St. Charles, Missouri. Des données détaillées de vitesse ont été rassemblées dans deux conditions différentes d'écoulement le long des formes du lit dans ce fleuve sableux. Les données à l'échelle des grands fleuves reflètent des caractéristiques d'écoulement semblables à celles des écoulements en laboratoire, avec la séparation d'écoulement se produisant en aval de la crête de dune et le recollement de l'écoulement sur la face de la dune suivante Des réponses de sillage de l'écoulement aux dunes ont été détectées, avec la vitesse diminuant dans tout le tirant d'eau au-dessus des creux, et la vitesse augmentant au dessus des crêtes. Les distributions des vitesses moyennes locales et spatiales étaient logarithmiques pour les deux ensembles de données. L'écart du profil des vitesses moyennes dans l'espace ramené au modèle standard de défaut de vitesse a été évalué. Les vitesses moyennes verticales ramenées à une moyenne d'écoulement pour le modèle de défaut de vitesse étaient à moins de 5% des valeurs mesurées et les vitesses moyennes dans l'espace étaient à moins de 10% pour les 90% du tirant d'eau supérieur. Le modèle de défaut de vitesse, négligeant la fonction de sillage, a été évalué pour estimer la vitesse moyenne sur la verticale à moins de 1% des valeurs mesurées.
In spite of significant developments made in turbulent boundary layer research over the years, our understanding of Ž. particle interactions with organized flow configurations coherent structures and their implications for mechanics of bedform development is far from completion. Researchers have made sincere efforts to obtain fundamental understanding of bedform mechanics through theoretical and experimental studies supplemented by field observations. A brief outline emphasising the nature of recent advancements and their consequences in sediment transport and bedform mechanics in aqueous environment under unidirectional current is given in this paper.
In river rehabilitation studies it is necessary to assess the impacts of the modified sand content, namely on near-bed flow hydrodynamics. This laboratory study addresses the changes that different bed morphology configurations and sediment transport impose on hydrodynamic variables relevant for the momentum and turbulent kinetic energy (TKE) budgets. Bed morphology was characterized by the void function, sand content, basal porosity and bed thickness. Special emphasis was given to the characterization of mean Reynolds and form-induced stresses, near-bed pressure, mean longitudinal velocity, momentum diffusivity and TKE production. Five laboratory tests simulated different stream conditions and sand contents in a framework-supported gravel bed. Instantaneous velocity maps were obtained with Particle Image Velocimetry (PIV). The collected data was analysed and theoretically framed with double-averaged methods (DAM). The impacts of the variations in the bed morphology over the near-bed flow hydrodynamics are discussed by direct comparison of the normalised DA flow variables. It was found that the effect of near-bed sediment movement is felt primarily in the wake production terms, in the shear rate below the plane of the crests and, to a lesser extent, in the form-induced shear stresses.
This paper uses particle imaging velocimetry to provide the first measurements detailing the flow field over a porous bed in the presence of bed forms. The results demonstrate that flow downstream of coarse-grained bed forms on permeable beds is fundamentally different to that over impermeable beds. Most significantly, the leeside flow separation cell is greatly modified by jets of fluid emerging from the subsurface, such that reattachment of the separated flow does not occur and the Reynolds stresses bounding the separation zone are substantially lessened. These results shed new light on the underlying flow physics and advance our understanding of both ecological and geomorphological processes associated with permeable bed forms. Water fluxes at the bed interface are critically important for biogeochemical cycling in all rivers, yet mass and momentum exchanges across the bed interface are not routinely incorporated into flow models. Our observations suggest that ignoring such exchange processes in coarse-grained rivers may overlook important implications. These new results also provide insight to explain the distinctive morphology of coarse-grained bed forms, the production of openwork textures in gravels, and the absence of ripples in coarse sands, all of which have implications for modeling and prediction of sediment entrainment and flow resistance.
Sediment transport processes in rivers continue to pose a challenge when designing movable-bed physical models, particularly for reproducing the grain sorting and bank erosion (fluvial erosion and mass failure). This paper presents and discusses scale effects of a specific scaling approach for multi-grain size mixtures that preserves similarity of initial motion for each grain size class and of the bank stability coefficient between the model and the prototype, but relaxes strict similarity of the Shields and particle Reynolds numbers. This approach is appropriate when bed load transport near incipient motion conditions is being studied, and allows for larger grain size scales than when full Shields parameter similarity is enforced. As part of an environmental project to rehabilitate sediment transport through bank erosion, this method has been applied to scale a Froude number criterion physical model of a reach of the Old Rhine (France). This has resulted in an undistorted scale of 40, and the use of sand as the model bank material. Each grain size has a different geometrical scale. The time scale for sediment motion is grain size and flow discharge dependent. An average time scale of 6 has therefore been used (four model hours ¼ one prototype day). A strategy devised for the field case consists of two higher, larger island groynes that replace the three existing groynes, producing bank erosion for flow rates below the mean annual flow rate. Extrapolation of model behaviour to the prototype is not a major problem, but the volume of eroded bank material may be underestimated, mainly because of the relaxation of the Shields number similarity and the apparent cohesive properties of the model bank material. Ó
Iranian Journal of Science & Technology, Transaction …, 2005
The transport of sediment in open channels is a complex process, and the physics of this phenomenon have not been completely explored. The majority of research work on sediment transport has been concentrated on beds formed of the same mobile sediment and only a few researchers have been concerned with sediment motion over fixed bed. This paper reviews the state of the problem and focuses on some practical points. Sediment threshold experiments were conducted in the two types of V-shaped bottom channels. Sand and gravel particle movements were considered and the relationship between flow discharge and bed shear stress, as well as channel bed slope were found at the threshold condition. Some practical and design equations were found to be more appropriate. It may be found that the effect of cross sectional shape on sediment threshold in fixed bed channels should be examined.