Open access peer-reviewed chapter

Physics of Cohesive Sediment Flocculation and Transport: State-of-the-Art Experimental and Numerical Techniques

Written By

Bernhard Vowinckel, Kunpeng Zhao, Leiping Ye, Andrew J. Manning, Tian-Jian Hsu, Eckart Meiburg and Bofeng Bai

Reviewed: 01 March 2022 Published: 03 May 2022

DOI: 10.5772/intechopen.104094

From the Edited Volume

Sediment Transport - Recent Advances

Edited by Andrew J. Manning

Chapter metrics overview

234 Chapter Downloads

View Full Metrics

Abstract

Due to climate change, sea level rise and anthropogenic development, coastal communities have been facing increasing threats from flooding, land loss, and deterioration of water quality, to name just a few. Most of these pressing problems are directly or indirectly associated with the transport of cohesive fine-grained sediments that form porous aggregates of particles, called flocs. Through their complex structures, flocs are vehicles for the transport of organic carbon, nutrients, and contaminants. Most coastal/estuarine models neglect the flocculation process, which poses a considerable limitation of their predictive capability. We describe a set of experimental and numerical tools that represent the state-of-the-art and can, if combined properly, yield answers to many of the aforementioned issues. In particular, we cover floc measurement techniques and strategies for grain-resolving simulations that can be used as an accurate and efficient means to generate highly-resolved data under idealized conditions. These data feed into continuum models in terms of population balance equations to describe the temporal evolution of flocs. The combined approach allows for a comprehensive investigation across the scales of individual particles, turbulence and the bottom boundary layer to gain a better understanding of the fundamental dynamics of flocculation and their impact on fine-grained sediment transport.

Keywords

  • cohesive sediment
  • floc measurement
  • particle-resolved direct numerical simulation
  • continuum model
  • population balance equation

1. Introduction

Cohesive sediment transport is a tightly coupled system driven by hydrodynamic forcing, resuspension, deposition, and flocculation. Turbulence intensity in the bottom boundary layers and the water column mixed layers controlled by currents (e.g., river outflows, tidal currents), surface waves, and estuarine circulations can vary by several orders of magnitude in terms of turbulent dissipation rate. Hence, the turbulent shear rate defined as G=ε/ν is the most important parameter for flocculation models. In the estuarine tidal boundary layers, the turbulent dissipation rate is around O106104m2/s3 (G=110s1; e.g., [1, 2]). Near estuarine fronts with shear instabilities, turbulent dissipation rates can increase up to O103m2/s3 (G=30s1; [3]). During storm condition when wave motion become dominant, turbulent dissipation rates in the thin wave bottom boundary layer can exceed O103 (G=30100s1, [4, 5]). Under breaking waves in the upper ocean or in the surf zone, the turbulent dissipation rate can be as high as O103102(G=100200s1, [6, 7]).

Computer simulation models are commonly the chosen tools that coastal managers use to predict sediment transport rates. In these continuum models, the fluid motion is computed via the continuity equation and the Navier-Stokes equations or spatially and temporally averaged variants thereof, whereas the sediment is represented as a concentration field [8]. This approach allows to simulate large spatial scales covering entire estuaries, because the governing equations are solved on the same Eulerian grid. However, these types of models require closures to account for the unresolved physics of the sediment dynamics, in particular vertical sedimentary distributions [9, 10] and mass fluxes. The latter is the product of the concentration and the settling velocity. Manning and Bass [11] found that mass settling fluxes can vary over four or five orders of magnitude during a tidal cycle in mesotidal and macrotidal estuaries; therefore, a realistic representation of flux variations is crucial to an accurate depositional model.

The specification of the flocculation term within numerical models depends on the sophistication of the model structure. Until recently, even the conceptual relationship between floc size, suspended particulate matter (SPM) concentration and turbulent shear stress proposed by Dyer [12] (see Figure 1a) remained largely unproven. Hence, much more work needs to be done in order to arrive at robust simulation tools with predictive capacity. The present contribution, therefore, provides a review on the state-of-the-art of floc measurements in both the field and laboratory. In addition, we review a newly emerging technique of particle-resolved simulations that can provide a promising alternative avenue to generate data of small-scale sediment dynamics to derive the missing constitutive equations for continuum models. Finally, we provide a summary of current techniques that are used in continuum models to account for cohesive sediment dynamics, where we explicitly point to the components that deserve more research in the future.

Figure 1.

An illustration of the dependence of floc size on shear stress and SPM concentration is given. (a) A conceptual diagram showing the relationship between floc modal diameter, suspended sediment concentration and shear stress (from [12]). Subplots (b–e): illustrative examples of real estuarine floc images showing ambient shear stress, SPM concentration and settling velocity; (b) shear stress 0.3 N m2, SPM concentration 3.5 g l1, and settling velocity 5 mm s1; (c) shear stress 0.3 N m2, SPM concentration 3.5 g l1, and settling velocity 8 mm s1; (d) shear stress 0.45 N m2, SPM concentration 0.25 g l1, and settling velocity 1.8 mm s1; (e) shear stress 0.45 N m2, SPM concentration 0.25 g l1, and settling velocity 0.2 mm s1. Illustrations (b–e) are modified from Manning.

Advertisement

2. Assessment of temporal floc evolution

2.1 Floc measurements

2.1.1 Direct floc size measurements

The presence of large estuarine macroflocs was initially observed in situ using underwater photography [13]. However, floc breakage occurs during sampling in response to the additional shear created by the instrumentation [14]. To overcome this problem, less-invasive techniques for measuring floc properties in situ have been developed. Usually, these can be divided into devices that solely measure floc size (D) e.g. Lasentec (Figure 2a) [15], LISST [16], LISST-Holo [17], and InSiPid [18]; and those devices that can provide measurements both of floc size and settling velocity (Ws) e.g. VIS [19], HR Wallingford video camera system [20], VIL [21], INSSEV: IN-Situ SEttling Velocity instrument [22, 23] (Figure 2bc), LabSFLOC—Laboratory Spectral Flocculation Characteristics—instrument [25, 26, 27], INSSEV-LF [28], and PICS—Particle Imaging Camera System [29].

Figure 2.

Some examples are given of floc measuring instrumentation. (a) Schematic of the adapted Lasentec par-tec 100 probe unit showing: 1. The light guide, scanning mechanism and focusing lens, 2. The PVC cylinder, 3. Watertight cable termination, and 4. Electronic circuitry and power supplies (from [15]). (b, c) the INSSEV instrument (from [24]); (b) side view of INSSEV mounted on a metal deployment frame; (c) front view of INSSEV (right side of image), together with optical backscatter (OBS) sensors and an acoustic Doppler velocimeter (ADV) positioned on a vertical pole (left side of image). The ADV provides high frequency turbulence data that can be directly related to the floc populations. (d) Views of a LabSFLOC-2 together with a schematic illustration of the instrument after [25]).

The strength of video-based floc measurements is that they minimize the number of assumptions used during the data processing and interpretation stages. Devices that only measure the size component require additional gross and often incorrect assumptions regarding the relationship between settling velocity, floc size, and floc density. The settling velocity of a floc is a function of both its size and effective density, and both of these floc components can display variations spanning three to four orders of magnitude within any one floc population [30, 31, 32].

Of note, the LabSFLOC suite of high-resolution, low intrusive, underwater video camera systems for the past 20+ years have been regarded internationally as a benchmark device for the sampling and dynamical simultaneous measurement of sizes and settling velocities for entire populations of flocs and a range of bio-sedimentary particles, and this enables individual floc effective density values to be determined by applying a modified Stokes Law [33]. This also enables the calculation of additional floc properties including: structural composition (porosity, fractal dimension), shape, sedimentary mass flux, and floc population mass-balancing, all from within a wide range of aquatic environments. In conclusion, selection of the most appropriate instrumentation is paramount when attempting to parameterise flocculated cohesive sediments. Manning et al. [34] provide a detailed review of many of these floc measuring systems.

2.1.2 Parameterizing floc data

To aid the interpretation of floc characteristics and for their inclusion in sediment transport models, each floc population can be segregated into various subgroupings according to floc size. Sample-mean floc values can be computed (i.e., a single value per floc population) to show generalized floc property trends.

Dyer et al. [35] reported that a single mean or median settling velocity did not adequately represent an entire floc spectrum, especially in considerations of a flux to the bed. Dyer et al. [35] recommended that the best approach for accurately representing the settling characteristics of a floc population was to split a floc distribution into two or more components, each with their own mean settling velocity. Both Eisma [13] and Manning [32] concur with this finding by suggesting that a more realistic and accurate generalization of floc behavior can be derived from the macrofloc and microfloc fractions. These two floc fractions form part of Krone’s [36] classic order-of-aggregation theory and produce two floc property values per floc population.

Macroflocs are large (typically D>160μm [32]) highly porous (typically >90%) and often fragile (and difficult to sample), fast-settling aggregates (see Figure 1bd), and typically close to the size of the turbulent Kolmogorov microscale [37, 38]. They are recognized as the most important subgroup of flocs because their fast-settling velocities tend to have the strongest influence on the mass settling flux [39]. Macroflocs are progressively broken down as they pass through regions of higher turbulent shear stress and reduced again to their component microfloc substructures [40]; they rapidly attain equilibrium with the local turbulent environment.

The smaller microflocs (typically D<160μm [32]; see Figure 1e) are considered to be the building blocks from which the macroflocs are comprised. Many field studies [22, 41, 42, 43] have shown that the microfloc class of aggregates tend to display a much wider range of effective densities and settling velocities than the macrofloc fraction. Microflocs are much more resistant to breakup by turbulent shear; they tend to have slower settling velocities but exhibit a much wider range of effective densities than the larger macroflocs.

In terms of flocculation kinetics [44], the macroflocs tend to control the fate of purely muddy sediments in an estuary [45]; this is because the smaller microflocs generally settle at less than 1 mm s1, whereas macroflocs settle in the 1–15 mm s1 range, enabling them to deposit to the bed [46]. However, when flocculation of mixed sediments occurs, the microflocs can potentially demonstrate settling velocities comparable to those of the macroflocs [28, 47].

2.1.3 Floc data example

In order to illustrate the spectral variability of floc properties, each floc population can be divided into various size bands. The band divisions can be chosen to best fit the data collected. LabSFLOC data for a mud sample from the Medway Estuary, UK, provide a graphical illustration of size banding and the increasing settling velocities associated with larger flocs (Figure 3). Twelve size bands (SB) have been used to represent the Medway floc data, with SB1 representing microflocs less than 40μm in size, whilst SB12 is representative of macroflocs greater than 640μm in diameter; SB2 to SB6 range from 40 to 240μm in five steps, each of 40μm, and SB7 to SB11 range from 240 to 640μmm in five steps of 80μm.

Figure 3.

An example of various size-banded floc properties for a LabSFLOC SPM sample is given. The complete population of size versus settling velocity data is illustrated in the top-left panel. These data apply to Medway estuary mud slurry (1.6 g l1) that has been sheared in the Southampton oceanography Centre (UK) mini-flume at a shear stress of 0.37 nm–2 (data from results in [43]).

2.1.4 Floc settling modeling approaches

2.1.4.1 Constant settling velocity

Specification of the flocculation term within numerical models depends on the sophistication of the model. The simplest parameterisation is a single floc settling velocity value that remains constant in both time and space (one coefficient). These fixed settling values are usually in the range of 0.5–5 mm s1 and typically are selected on an arbitrary basis and sometimes used as a tuning parameter to match predicted erosion and deposition patterns to observations for an undisturbed estuary.

2.1.4.2 Power law settling velocity

The next step is to use gravimetric data provided by field settling-tube experiments to relate floc settling velocity to the instantaneous SPM concentration, using a power law with two coefficients (e.g., [48]; see Figure 4a). Empirical results have shown a generally exponential relationship between the mean, or median, floc settling velocity and SPM concentrations for concentrations <10gl1. This approach sometimes includes hindered settling (see Figure 4b). However, both the constant settling velocity and the power law parameterisation techniques do not include the important and influential effects of turbulence such as floc breakup induced by turbulent shear.

Figure 4.

Some examples of floc settling velocity measurements are shown. (a) Owen tube determinations of median settling velocity as a function of suspended sediment concentration for different estuaries; the bold dashed line represents an exponent of unity (reproduced with minor modifications from [49]). (b) Median settling velocity of Severn estuary mud as a function of SPM concentration; the Owen-tube data are taken from Odd and Roger [50], and the dashed line represents results based on the SandCalc software sediment-transport computational algorithm, which incorporates the hindered settling effect at high concentrations (reproduced with minor modifications from Soulsby [51]).

2.1.4.3 The van Leussen parameterisation

More recently, a number of authors have proposed simple theoretical formulae interrelating a number of floc characteristics that can then be calibrated using empirical studies. Such an approach has been used by van Leussen [19], who utilized a formula that modifies the floc settling velocity in still water by a floc growth factor, due to turbulence, and then reduces it by a turbulent floc disruption factor. The reference settling velocity (taken at low turbulent shear conditions), Ws0, is then related to the SPM concentration (C) by a power law:

Ws0=kCmE1

where k and m are empirical constants. The van Leussen is a qualitative simplification of a model originally developed for the sanitation industry [52], with only a limited number of interrelated parameters, and hence does not provide a complete description of floc characteristics within a particular sheared environment.

2.1.4.4 The Lick et al. parameterisation

A number of authors have attempted to observe how the floc diameter changes in turbulent environments. In particular, Lick et al. [53] derived an empirical relationship based on laboratory measurements made in a flocculator. They found that the floc diameter varied as a function of the product of the SPM concentration and the turbulence parameter as the turbulent shear rate, G:

D=cCGdE2

where c and d are empirically determined values. However, this formulation provides no information on the important floc settling or floc dry mass properties.

2.1.4.5 The Manning and Dyer parameterization

The Manning Floc Settling Velocity (MFSV) algorithm for settling velocity [54] is based entirely on empirical observations made in situ using nonintrusive floc and turbulence data acquisition techniques in a wide range of estuarine conditions. The floc population size and settling velocity spectra were sampled using the video-based INSSEV instrument and LabSFLOC data.

The Manning-Dyer algorithms were generated by a parametric multiple regression statistical analysis of key parameters, which were generated from the raw, spectral floc data. Detailed derivations and preliminary testing of the floc-settling algorithms are described by Manning [26, 55]. Although the resulting empirical formulae are not presented in a fully dimensionless form, these formulae have the merit of being based on a large dataset of accurate, in situ settling velocity measurements (157 individually observed floc populations), acquired from different estuaries (Tamar, Gironde and Dollard) and different estuarine locations, such as the turbidity maximum and the intertidal zone.

The algorithms are based on the segregation of flocs into macroflocs (D>160μm [32]) and microflocs (D<160μm), which comprise the constituent particles of the macroflocs. This distinction permits the discrete computation of the mass settling flux (MSF) at any point in an estuarine water column.

Equations are given for: the settling velocity of the macrofloc fraction (Ws,macro), the settling velocity of microflocs (Ws) and the ratio of macrofloc mass to microfloc mass in each floc population—termed the SPM ratio [55]. This type of formulation gives a good compromise between the representation of physicochemical processes and computational simplicity. Eqs (3) and (4) describe Ws,macro (mms1) with inputs of SPM in mgl1 and τ in Pa:

Ws,macro=0.644+0.000471SPM+9.36τ13.1τ2E3

for 0.04<τ<0.6 Pa

Ws,macro=3.96+0.000346SPM4.38τ+1.33τ2E4

for 0.6<τ<1.5 Pa. These equations require the input of a turbulent shear stress (τ) and an SPM concentration. These regression equations provide a realistic approximation to the field data. Graphical representations of the equations, together with the data, are presented in Manning and Dyer ([54]; see Figure 5). The Manning settling algorithm is valid for SPM concentrations in the range 10–8600 mg l1 and shear stress values of τ<2.13 Pa, with extrapolation extending this range to 5–10 Pa.

Figure 5.

Illustration of the settling velocities of macroflocs and microflocs, plotted against shear stress, for a mixed sediment suspension comprising a ratio of 25 per cent mud to 75 per cent sand and a pure mud suspension, all for a total SPM concentration of 5 gl–1 (modified from Figure 14 of Manning et al. [56]).

An example of this is the implementation of the algorithm in a TELEMAC-3D numerical model of the Thames Estuary, UK [57], in which it was shown that the use of the Manning algorithm greatly improved the reproduction of observed distributions of SPM concentrations compared with the other formulations, both in the vertical and horizontal dimensions.

The Manning settling algorithms have been extended to cater for mixed sediment flocculation settling, including different ratios of mud:sand ([24, 47, 56]; see Figure 5). These algorithms are a major step forward in establishing a reliable estimate of the settling velocity. It has been developed based on a large and reliable dataset, it caters for the spectrum of hydrodynamic conditions that occur during a typical tidal cycle [58] (a feature often lacking in the settling terms of many estuarine sediment transport models) and has been shown to more accurately reproduce the distribution of suspended sediment compared with simpler settling models.

Soulsby et al. [59] has developed a more ‘physics-based’ version of the empirical model based on the Manning-Dyer formulation, called Soulsby-Manning 2013. It should be noted that for flocculation algorithms and models that include turbulence as a contributing variable, it is vital to ensure that the turbulence data are accurate, otherwise it has significant implications for the accuracy of the calculated floc settling characteristics.

2.1.4.6 Complex population approaches

Lee et al. [60] applied a time-evolving two-class population balance equation (PBE) to determine the spatially and temporally changing distribution of fixed-size microflocs and size-varying macroflocs for bimodal floc distributions, with a fractal relationship between floc size and mass to derive the distribution of settling velocities. However, the authors felt that further intensive investigation of the aggregation and breakage kinetics would be required before their model was generally applicable when compared with the simpler approach of Manning and Dyer [54] and, presumably, Soulsby et al. [59].

Verney et al. [61] applied a time-evolving, multi-fraction model to determine the spatially and temporally changing distribution of the numbers of flocs in each size fraction, with a fractal relationship between floc size and mass to derive the distribution of settling velocities.

A relationship between the floc settling velocity and floc properties and fractal dimensions is given by Winterwerp [62]. A fractal approach has been used by Winterwerp [63] to solve a differential equation that simulates the time-varying representative floc diameter, from which floc density is derived from fractal considerations, and settling velocity obtained from a Stokes-like formula. Winterwerp et al. [64] also used a simplified fractal model to relate settling velocity to a turbulent shear parameter, the instantaneous concentration, and water depth.

The state-of-the-art model for floc structure is to assume a fractal structure. Many studies [65, 66, 67] indicate that the most sensitive parameters in a fractal model controlling the resulting settling velocity are the primary particle properties (primary particle diameter, density and their distributions) and the fractal dimension. For organic-rich particles, evidence suggests that the fractal dimension highly depends on stickiness [68].

Floc breakup in the existing size-class-based PBE formulation is modeled by assuming an invariant floc structure (e.g., fractal dimension 2) and key properties, such as floc yield strength, are assumed constant over a wide range of floc size. The importance of modeling floc yield strength as a function of floc size has been demonstrated by Son and Hsu [67] via a fractal concept [69, 70]. Further details are provided in Section 3.1 below.

A fractal model is widely used to parameterize floc structure. Fractals are a mathematical simple approach, where scientists feel computationally ‘comfortable’ and therefore are happy to ‘shoehorn’ all flocs into this framework, and a fractal dimension of around 2 is often used. However, its applicability for heterogeneous sediments remains to be proved. Moreover, for flocs of high organic content, stickiness can be significantly enhanced due to the presence of extracellular polymeric substances (EPS) and transparent exopolymer particles (TEP). Field observations suggest that the fractal dimension for inorganic particles is larger than 2.0 while for organic rich flocs, it can be smaller than 2.0 [68]. There are also empirical formulations suggesting that the fractal dimension depends on floc size [65, 66].

Unlike Verney et al. [61], who use floc diameter, Maggi et al. [71] describe the floc population based on the number of primary particles in the flocs, which appears to make the incorporation of a variable fractal dimension straightforward. Moreover, Maggi et al. [71] adopt a sophisticated collisional efficiency closure that considers the effects of floc size and permeability.

2.1.5 Future floc modeling directions

Real flocs are multi component of different densities; even measuring real fractal dimensions is highly problematic. The emergence theory has been utilized in many disciplines (e.g. [72, 73, 74]) and provides a valid alternative and potentially more realistic approach for representing real multi-component mud floc structures. Both Cranford et al. [75], and Rietkerk and van de Koppel [76] have successfully adapted an emergence approach for application to natural biomaterials and ecosystems (respectively).

By utilizing an emergence framework for flocculation, at one end a simple fractal representation would still operate for basic, geometrically repeating simple floc structures (e.g. flocs composed from a single clay primary particle). Whilst as the flocs before more complex in structure, composition and geometry, and fractal theory become less representative, the emergence would adapt to a more suitable floc representation. It is envisaged that this new emergence approach [77] could cope better and more efficiently and realistically for real flocs at a wide range in resolution scales all using real image data at each scale. This will provide a level of error checks that are not supported by regular fractal approaches. Nonetheless, we are some way off from implementing this approach in a numerical floc model and more fundamental research on floc dynamics, properties and characteristics is required, in particular 2D and 3D floc imaging techniques (e.g. [27, 78, 79, 80, 81, 82]).

2.2 Grain-resolving simulations

Grain-resolving simulations are a powerful tool to obtain detailed, high-fidelity data of complex fluid-particle systems. Despite its rather large computational costs, recent advances in computational power have made it possible to perform grain-resolving simulations on scales that become relevant for sediment transport phenomena. The idea is to compute the trajectories of all the individual grains in a flow. Typically, the flow is computed on a Eulerian grid that is fixed in space and time. It can either be approximated by assuming a prescribed background flow or by solving the Navier–Stokes equations. The movement of the particles is then computed by their equations of motion in a Lagrangian sense, i.e. the particle is free to move within the entire computational domain. Hence, this representation is commonly referred to as the Euler–Lagrange approach.

Depending on their fluid-particle coupling, several schemes can be employed. If the particles are driven by a fluid flow but do not modify the flow field, the scheme is considered one-way coupled, whereas the fluid-particle mixture is two-way coupled if the flow is modified by the particle motion as well. In addition, momentum exchange of particles may be accounted for by means of collision and contact or any other particle related force. Using a two-way coupled simulation that accounts for particle-particle interactions is considered a fully-coupled scheme. In the context of the present study, attractive cohesive forces can contribute to the particle-particle interaction by binding grains into aggregates that are much larger than the individual primary particle. This process can be understood as flocculation. In the following, we will first review the governing equations to compute the particle motion in a Lagrangian sense and then proceed to present two examples to model flocculation of cohesive sediment to investigate aggregation and settling processes.

2.2.1 Computing the particle motion

Regardless of the background flow, we prescribe the motion of each primary cohesive particle i as a sphere moving with translational velocity up,i and angular velocity ωp,i. These are obtained from the Newton-Euler equations

mp,idup,idt=Γp,iτndAFh,i+πDp,i3ρp,iρf6gFg,i+j=1,jiNpFcon,ij+Flub,ij+Fcoh,ijFc,i,E5
Ip,idωp,idt=Γp,ir×τndATh,i+j=1,jiNpTcon,ij+Tlub,ijTc,i,E6

where the primary particle i moves in response to the hydrodynamic force Fh,i, the gravitational force Fg,i and the particle-particle interaction force Fc,i which accounts for the direct contact force Fcon,ij in both the normal and tangential direction, as well as for short-range normal and tangential forces due to lubrication Flub,ij and cohesion Fcoh,ij, where the subscript ij indicates the interaction between particles i and j. The hydrodynamic torque is denoted by Th,i, while Tc,i represents the torque due to particle-particle interactions, where we distinguish between direct contact torque Tcon,ij and lubrication torque Tlub,ij. Here, mp,i denotes the particle mass, Dp,i the particle diameter, ρp,i the particle density, ρf the fluid density, g the gravitational acceleration, Γp,i the fluid–particle interface, τ the hydrodynamic stress tensor, Np the total number of particles in the flow and Ip,i=πρp,iDp,i5/60 the moment of inertia of the particle. Furthermore, the vector n represents the outward-pointing normal on the interface Γp,i, r is the position vector of the surface point with respect to the centre of mass of the particle.

Following Biegert et al. [83, 84], Zhao et al. [85, 86] represent the direct contact force Fcon,ij by means of spring-dashpot functions, while the lubrication force Flub,ij is accounted for based on lubrication theory [87] as implemented in Zhao et al. [85, 86]. The model for the cohesive force Fcoh,ij is based on the work of [88], which assumes a parabolic force profile, distributed over a thin shell of thickness hco surrounding each particle:

F˜coh,ij=4Coζn,ij2hcoζn,ijhco2n,ζmin<ζn,ijhco,0,otherwise.E7

Here, ζn is the gap size in between particle surfaces, ζmin is a limiter and Co is the cohesive number that has to be defined according to a given problem as will be detailed below. We remark that, based on Eqs. (5) and (6), the configuration of the primary particles within a floc can change with time in response to fluid forces, since the cohesive bonds are not rigid.

2.2.2 Aggregation

The flocculation process is strongly affected by the turbulent nature of the underlying fluid flow. Small-scale eddies modify the collision dynamics of the primary particles and hence the growth rate of the flocs, while turbulent stresses can result in the deformation and breakup of larger cohesive flocs. Hence, the dynamic equilibrium between floc aggregation and breakage is governed by a complex and delicate balance of hydrodynamic and inter-particle forces.

In the spirit of earlier investigations [89, 90], Zhao et al. [85] apply a simple model flow in order to investigate the effects of turbulence on the dynamics of cohesive particles. These authors consider the one-way coupled motion of small spherical particles in the two-dimensional, steady, spatially periodic cellular flow field commonly employed as initial condition for simulating Taylor-Green vortices (cf. Figure 6a), with fluid velocity field uf=ufvfT

Figure 6.

(a) Streamlines of the doubly periodic background flow given by Eq. (8); (b) typical floc configuration with individual flocs distinguished by color (figure taken from [85].

uf=U0πsinπxL0cosπyL0,vf=U0πcosπxL0sinπyL0,E8

where L0 and U0 represent the characteristic length and velocity scales of the vortex flow. By computing the fluid flow via this idealized flow field, the hydrodynamic torque Th,i can be omitted while the hydrodynamic force Fh,i is generally replaced by a simple stokes drag force Fd,i=3πDp,iμup,iuf,i when simulations are one-way coupled in the sense that the particles do not modify the fluid flow. Here, uf,i indicates the fluid velocity evaluated at the particle center, μ the dynamic viscosity of the fluid.

The dynamics of the primary particles are characterized by the Stokes number St=U0ρp,iDp,i2/18L0μ, and the settling velocity Ws=vs/U0 where vs=ρp,iρfDp,i2g/18μ is the Stokes settling velocity of an individual primary particle i, as well as the cohesive number

Co=maxFcoh,ijU02L02ρf=AHDp,i+Dp,j32λζ01U02L02ρf,E9

where the Hamaker constant AH is a function of the particle and fluid properties, λ=Dp,i+Dp,j/40 represents the range of the cohesive force and ζ0=Dp,i+Dp,j/8000 the characteristic distance. Representative values of AH for common natural systems can be found in [88].

Zhao et al. [85] employ a computational domain with periodic boundaries. All particles have identical diameters Dp and densities ρp. Initially they are at rest and separated, and randomly distributed throughout the domain. When the distance between two particles is less than λ/2, the particles are considered as part of the same floc and the number of flocs Nf is tracked as a function of time, with an individual particle representing the smallest possible floc (Figure 6b). Based on the simulation results, Zhao et al. [85] propose a new flocculation model to predict the temporal evolution of the floc size. For flocs of fractal dimension nf, the mean floc size D¯f is related to the average number of primary particles per floc N¯p,local=Np/Nf,

D¯f=N¯p,local1nfDp,E10
N¯p,local=11/N¯p,local,int1/N¯p,local,maxebt+1/N¯p,local,max,E11

where N¯p,local,int denotes the initial number of particles per floc and the average number of particles per floc during the equilibrium stage N¯p,local,max is defined as,

N¯p,local,max=Np,N¯p,local,maxNp,8.5a1St0.65Co0.58Dp2.9ϕ0.39ρs0.49Ws+10.38,otherwise,E12

where ρs=ρp/ρf denotes the density ratio, ϕ the volume fraction of particles. The agglomeration rate b with the constraint b0 is obtained by

b=0.7a2St0.36Co0.017Dp0.36ϕ0.75ρs0.11Ws+11.4,St0.7,0.3a2St0.38Co0.0022Dp0.61ϕ0.67ρs0.033Ws+10.46,St>0.7.E13

For the present cellular model flow the values a1=a2=1 in Eqs. (12) and (13) yield optimal agreement with the simulation data with the fitting deviation of ±30%. For real turbulent flows, a1 and a2 need to be determined by calibrating with experimental data. The new model, Eqs. (10)(13), with a constant fractal dimension nf=2 for predicting the floc size has been employed and successfully validated with experimental data in our earlier work [85] (cf. Figure 7). The recent study by Zhao et al. [86] even goes beyond this assumption by showing that the fractal dimension becomes a function of the floc size when particles undergo flocculation in isotropic turbulence.

Figure 7.

Calibration of the empirical coefficients for the models of Winterwerp [62] (kA=1.35 and kB=1.29×105), Kuprenas et al. [91] (kA=0.45 and kB=1.16×106), and for our Eqs. (10)(13) (a1=500 and a2=35); comparison between experimental data and predictions by the models. The experimental parameters are measured by Tran et al. [92], Dp=5μm, ρp=2650kg/m3, ρf=1000kg/m3, μ=0.001Ns/m2, the shear rate G=50s1, concentration C=200mg/L.

2.2.3 Hindered settling

It is well known that the settling behavior of a dense suspension differs substantially from the settling behavior of a single grain. Particles settling in a dense suspension induce a counterflow and experience friction by colliding with other particles. These processes yield the so-called hindered settling, which is substantially slower than the settling of an individual particle and depends on fluid and particle properties. Nevertheless, the Stokes settling velocity of an individual grain is still widely used to quantify the settling speed of sediment in particle-laden turbidity currents (e.g., [84]). Hence, constitutive equations to predict the settling speed as a function of the local flow conditions can enhance existing computational frameworks for the analysis of turbidity currents. To investigate the effects of the settling behavior of flocculating cohesive sediment by means of grain-resolving simulations [88, 93], it is important to not only account for frictional contact between particles in a dense suspension but also for the modifications of the fluid flow that is caused by settling particles displacing the fluid underneath them [94]. In this case, one needs to solve the Navier-Stokes equations for an incompressible Newtonian fluid:

ut+uu=1ρfp+νf2u+fIBM,E14

along with the continuity equation

u=0,E15

where u=uvwT designates the fluid velocity vector in Cartesian components, p denotes the pressure, νf is the kinematic viscosity, t the time, and fIBM represents an artificial volume force introduced by the Immersed Boundary Method (IBM) [95, 96]. This volume force is introduced in the vicinity of the inter-phase boundaries to enforce a no-slip condition on the particle surface and to modify the fluid motion according to the particle motion. This measure also yields the hydrodynamic forces and torques, Fh,i and Th,i in Eqs. (5) and (6), respectively, as a direct result of this coupling scheme.

While this set of equations represents a fully coupled system, it is important to note that the relevant scales that define the system have changed compared to Section 2.2.2. In the scenario that investigates hindered settling of polydisperse cohesive sediment, the fluid flow is driven by moving particles. Hence, the relevant scale becomes m50g, where m50 is the mass of the median grain size and g=ρp/ρf1g is the specific gravity of the sediment. This scaling yields a modified cohesive number:

Co=maxFcoh,ijm50gE16

and a characteristic particle Reynolds numberRe=D50us/νf, where D50 is the median grain size and us=gD50 is the buoyancy velocity of the sediment. These two non-dimensional numbers, Co and Re, fully define the physical system under consideration.

Particles that are placed in a tank will settle due to gravity thereby displacing the fluid and accumulate at the bottom of this tank. The induced counter flow as well as the particle-particle interaction yields frictional contacts and flocculation due to cohesive forces, which is the desired situation for hindered settling [88]. For small particle sizes, where cohesive forces remain relevant, Vowinckel et al. [88, 93] obtain a faster settling behavior for cohesive sediment as compared to its non-cohesive counterpart (Figure 8). During the settling process, the sediment will transform from a suspended state, where the weight is fully supported by the fluid pressure, to a deposited state, where the weight is supported through contact chains of the deposited sediment that extent all the way to the bottom of the tank [93]. This process is described by the effective stress concept, which states that the total stress, i.e. the submerged weight of sediment, is supported by either the particle pressure or the effective stress due to particle contact [97].

Figure 8.

Particle configurations during the settling and deposition process. Top row: Non-cohesive (Co=0)and bottom row: Cohesive (Co=5 according to Eq. (16)). Left column: t=17.6τs, which corresponds to the time at which the particle phase has its maximum kinetic energy. Here, τs=D/us is the non-dimensional reference time. From left to right, the columns are separated by time intervals of 72.5τs. The gray shading reflects the vertical particle velocity (figure taken from [88]).

Advertisement

3. Coastal modeling

Coastal modeling typically refers to the modeling of regional scale (10–100 km2) coastal and estuarine processes over a timescale of hours to months. Due to the large spatial and temporal scales that need to be covered, a Reynolds-averaged Navier-Stokes (RANS) model is adopted (e.g., [98, 99, 100]). Turbulence dissipation and mixing are parameterized with two-equation closure models via a diffusion process. Moreover, when surface waves are present, the individual wave-phase is often not resolved. The generation, transformation and dissipation of the random wave field are represented by a wave spectrum and solved by using the spectral wave action balance equation [101]. Wave-period-averaged wave statistics are then coupled with the coastal models. Consequently, the wave bottom boundary layer processes cannot be directly resolved and additional parameterizations are needed, such as the apparent roughness [102], i.e. the effect of the wave bottom boundary layer on the current resolved by the coastal model, wave-driven small-scale seabed morphological features (e.g., ripples), and the near-bed sediment transport processes (often called bedload or near-bed load). The suspended load transport for a range of non-cohesive sediment classes above the wave bottom boundary layer can be resolved in the coastal models via the conservation of mass. One of the main challenges for extending these coastal models for simulating cohesive sediment transport is the parameterization of settling velocity due to flocculation.

Since the recognition of flocculation in controlling the settling velocity of cohesive sediment in coastal and marine environments [12, 41], significant progress has been made, particularly in the physical parameters controlling the floc dynamics, floc size distribution and their relationships with settling velocity statistics. To name a few, the role of turbulent shear and particle concentration in determining the aggregation rate and the resulting floc size has been quantified [62, 103, 104, 105] (see also Figure 1a). Particularly, in many tidal boundary layers and laboratory experiments of homogeneous turbulence, the mean floc size is observed to be limited by the Kolmogorov length scale (e.g., [106, 107]). Moreover, the relationship between floc size population and settling velocity (or floc density) in both idealized and realistic conditions has been revealed (e.g., [108, 109]) and the fractal dimension has been applied to model these relationships [65, 66, 69, 71] (see also Figure 3). Finally, researchers have begun to understand complex floc characteristics in estuaries dominated by organic particles (e.g., [110]), high cohesion due to TEP (e.g., [111]) and the presence of sand (e.g., [112, 113], see also Figure 5). A more complete discussion on these observational-based empirical parameterizations is provided in Section 2.1.4.

Advancements have also been made in modeling flocculation processes of cohesive sediments. Following the summary presented in Section 2.1, this section focuses on a more in-depth discussion of the complex population approach. Pioneering work by Winterwerp [62, 94] established a robust single-size class (averaged floc size) flocculation modeling framework. This framework has been refined by Kuprenas et al. [91] to limit the floc size growth by the Kolmogorov length scale. A more sophisticated flocculation model [114] based on the population balance equations (PBE) has been incorporated into the Princeton Ocean Models (POM) by Xu et al. [115] to study the dynamics of Estuarine Turbidity Maximum (ETM). Recently, Sherwood et al. [116] incorporated the PBE flocculation model FLOCMOD by Verney et al. [2] into the Regional Ocean Modeling system (ROMS), which is part of the Coupled Ocean-Atmosphere-Wave-Sediment Transport Modeling System (COAWST, [98]). The model is used to study cohesive sediment transport in an idealized setting and a realistic application in the York River estuary. Through a direct simplification of the PBE type model, a tri-modal flocculation model was recently developed in the coastal model TELEMAC [117]. Last but not least, the empirical parameterization of the floc settling velocity MFSV algorithm (see Section 2.1.4) suggested by Manning and Dyer [54] has been incorporated into TELEMAC-3D [57], while more recently the nondimensional version of MFSV proposed by Soulsby et al. [59] has been incorporated into the Finite Volume Community Ocean Model (FVCOM, [99]). These closure models cover a wide range of flocculation physics incorporated or neglected, and the resulting computational cost also varies from minimal to very significant. It is worth to point out that besides the turbulence-averaged models discussed so far, a PBE formulation for flocculation dynamics has recently been incorporated into a turbulence-resolving large-eddy simulation model by Liu et al. [118] to study floc dynamics in the upper-ocean mixing layer subject to Langmuir turbulence.

3.1 Structure of continuum model

The Reynolds-averaged suspended sediment mass concentration Cxit is solved by the conservation of mass in an advection-diffusion equation:

Ct+uiCxi=CWsδi2xi+xiν+νtCxiE17

where ui is the fluid velocity, Ws is the settling velocity, ν is the fluid viscosity and νt is the turbulent (eddy) viscosity. The Kronecker delta δij is used here with j=2 representing the direction of gravitational acceleration. A challenge in modeling cohesive sediments is that the settling velocity, Ws, depends on the flocculation process. To include flocculation, there are generally two approaches [119]. The first approach is called the distribution based approach (e.g., [116, 120]), which directly models a bulk settling velocity as a variable due to flocculation. The simplest formulations are empirical formulas characterizing settling velocity as a function of mean floc size, sediment concentration, and turbulent shear rate [64]. Most notably, the framework provided by Winterwerp [62, 94] solves flocculation using a mean floc size, or number concentration of floc (having a mean floc size) that explicitly includes aggregation and breakup terms. Floc density and settling velocity can then be estimated by assuming a constant fractal dimension nf. Such framework has been extended for a variable fractal dimension [67] and applied to model sediment resuspension in the Ems/Dollard estuary by Son and Hsu [121]. Flocculation process involved complex interaction between floc aggregation and breakup of different sizes, ranging from primary particles, microflocs, and macroflocs (see Section 2.1.2 for their definition). Using a mean value of floc property to describe these complex flocculation process may be too simplistic. For example, field observations suggest that natural flocs sometimes show a bimodal distribution (e.g., [18]). Maerz et al. [120] developed a flocculation model that solves the first moment of floc size distribution. Shen and Maa [119, 122] further use the quadrature of moments to model the evolution of floc size distribution. As discussed by Sherwood et al. [116], since settling velocity calculated from the distribution-based approach are based on the statistics of floc properties, the resulting bulk settling velocity must be a variable in time and space. Many existing coastal models utilize an efficient advection scheme that is designed for a constant and uniform settling velocity [115, 123]. Since these coastal models have been designed and routinely used to model suspended sediment (non-cohesive) transport of multiple size classes, it is more straightforward to extend the multiple-size class model into a PBE formulation to model flocculation processes.

3.2 Population balance equation

In the population balance formation, the sediment mass concentration C is partitioned into N classes and each class is represented by an index k, where k and N are positive integers. N must be a sufficiently large number to resolve the distribution. The sediment mass concentration in each class ckxit is calculated by its mass conservation equation:

ckt+uickxi=ckws,kδi2xi+xiν+νtckxi+G1,kL1,k+,k=1,2,NE18

where ws,k is the settling velocity of class k. To include the flocculation processes, additional gain and loss terms, e.g., G1,k,L1,k,, due to aggregation of flocs (and primary particles) and breakup of flocs need to be modeled, which will be discussed later. It is important to point out that the flocculation processes only re-distribute the floc mass among different size classes and care must be taken to ensure the total sediment mass conservation when modeling these terms (i.e., these gain and loss terms must cancel each other after the summation of all size classes in Eq. (18)), including the numerical treatment (e.g., using a logarithmically distributed size class and a mass-weighted interpolation).

As mentioned before, the sediment mass concentration distribution can be described by floc size class [61] or number of primary particles in the floc [66]. Here, we focus on the more popular one using floc size class with each class having a floc diameter of Df,k. The sediment mass concentration of each class can be related to number concentration of flocs in each size class nkxit as:

ckxit=mknkxitE19

where mk is the mass of a floc in size class k. Assuming a floc is formed by spherical primary particles of diameter Dp and density ρs, the mass of a primary particle can be calculated as

mp=π6ρsDp3.E20

By, furthermore, assuming that the floc structure follows a fractal relationship, we can calculate the mass of a floc as [69]:

mk=mpDf,kDpnf.E21

Therefore, with the given fractal dimension and primary particle properties, mk can be explicitly calculated for a given floc size class Df,k. Following the fractal theory, the floc density of size class k can be readily calculated as [69].

ρkf=ρw+ρsρwDf,kDpnfE22

where ρw is the density of water (or seawater). With the known floc diameter and floc density, the settling velocity of the flocs in each size class can be calculated. The simple Stokes law is used here following Sherwood et al. [116]

ws,k=ρkfρwgDf,k218μE23

where μ is the dynamic viscosity of water (seawater).

It is worth noting that in the population balance formulation, the floc settling velocity of a particular size class is treated as a constant determined by the given fractal dimension, primary particle properties, and fluid properties. Therefore, it is more suitable for typical coastal modeling systems due to their numerical treatment of advection. In typical field conditions, a size-class based population balance formulation is reported to require at least 10–20 size classes [61, 116].

After proposing the appropriate gain and loss terms in Eq. (18), the full dynamics of floc transport, settling and re-distribution of sediment mass among all floc size classes due to flocculation can be modeled. In practice, some models solve a system of partial differential equation for the number concentration nkxit directly by substituting Eq. (19) into Eq. (18) (e.g., Liu et al. [118]). For typical coastal models (e.g., Sherwood et al. [116]), the numerical treatment of flocculation and the advection-diffusion-settling processes are split into two steps. The zero-dimensional (homogenous turbulence condition without the advection, diffusion and settling terms) number concentration equations that only include the gain and loss terms are solved at every grid point over each (baroclinic) time step size of the coastal model to redistribute floc mass between different classes. Then, the newly calculated floc mass due to flocculation in each size class is updated by the advection-diffusion-settling equation (Eq. (18)) without the gain and loss (flocculation) terms. One advantage of such approach is that many zero-dimensional flocculation models developed elsewhere can be easily coupled into the coastal models. For example, Sherwood et al. [116] couple an existing zero-dimensional size-class based flocculation model FLOCMOD developed by Verney et al. [61] into the COAWST coastal modeling framework.

In this paper, the essential model formulations and closures of Verney et al. [61] are reviewed. The purpose here is not to discuss the model details, since they can be found in the cited references. Rather, this section is intended to bridge the discussions in Section 1 and Section 2 by focusing on the key model elements that are sensitive to the model results and hence may require more physical understanding. The governing equation for floc number concentration nk (S.I. unit: m3) of size class k is written as

nkt=GakLak+GbskLbskE24

in which Gak and Lak represent the gain and loss of flocs in size class k due to aggregation, while Gbsk and Lbsk represent the corresponding gain and loss due to shear-driven breakup. Verney et al. [61] also include terms due to collision-induced breakup [70]. However, this mechanism is shown to be of minor importance for pure clay flocculation and it is not discussed here for brevity. The model assumes aggregation that is driven by turbulent shear and binary collision. Other known mechanisms driving aggregation, namely, the Brownian motion (important for particle smaller than 1μm) and the differential settling are neglected by assuming that coastal and estuarine environments are dominanted by turbulent shear. The Gak and Lak terms are modeled as

Gak=12i+j=kαijAijninjE25

and

Lak=iNαikAikninkE26

where αij represents the collisional efficiency (dimensionless) and the shear-driven binary collision probability function is written as

Aij=16GDi+Dj3.E27

The quantity Gak shown in Eq. (25) represents the gain of flocs in size k due to aggregation of smaller flocs at size class i and j=ki while Lak shown in Eq. (26) expresses the loss of flocs in size class k due to aggregation with flocs in other size classes. In Verney et al. [61], αij is set to be a constant αij=α for simplicity. This point will be revisited later. Gain and loss of flocs in size class k due to breakup driven by turbulent shear are calculated as

Gbsk=i=k+1NΠkiBiniE28

and

Lbsk=Bknk.E29

Essentially, gain of flocs Gbsk in size class k due to fragmentation can only occur in floc size classes larger than k. The fragmentation probability function is written as

Bi=βiG3/2DiDiDpDp3nfE30

where βi is the (dimensional: m1s1/2) fragmentation rate and it is assumed to be a constant βi=β in Verney et al. [61]. More discussions on βi will be given later. The breakup distribution function is defined as Πki and it characterizes different breakup scenarios. Our understanding on the shear-induced breakup scenarios is currently limited. Nevertheless, the breakup distribution functions Πki provided in Verney et al. [61] include binary (fragmentation of floc mass mk into two small flocs of equal mass mk/2), ternary (fragmentation of floc mass mk into three smaller flocs, one floc of mk/2 and two flocs of mk/4), and erosion (fragmentation of floc mass mk into r+1 flocs, one larger floc and r smaller flocs of equal size).

As demonstrated in Verney et al. [61], their PBE-based flocculation model can predict several key features of floc dynamics observed in the field. For instance, the model is capable of reproducing the observed slower aggregation and more rapid fragmentation process (so-called clock-wise hysteresis of aggregation/fragmentation process) during a tidal cycle. As the floc size directly controls the settling velocity, capturing the hysteresis of floc aggregation/fragmentation is essential to further predict the net sediment fluxes during a tidal cycle. Moreover, Verney et al. [61] showed that a bimodal distribution of flocs often observed in the field can be reproduced by the PBE model by including a mix of different breakup distribution functions. Although the PBE-based flocculation models provide a promising modeling framework for cohesive sediment transport, there are limitations that require future investigations.

The most sensitive empirical parameters in FLOCMOD are the collisional efficiency α and the fragmentation rate β, which are often assumed to be constants. As demonstrated systematically by Verney et al. [61], to match the mean equilibrium floc size, there is no unique set of optimum α and β values and it is the ratio of α/β for these two empirical parameters that controls the mean equilibrium floc size. A similar finding is reported in the single size-class flocculation model of Winterwerp [62]. Hence, to fully benefit from the capability of PBE-based models that provide the temporal evolution of the full spectrum of floc sizes, the model calibration should go beyond just using equilibrium mean floc size. Sherwood et al. [116] provide a limited validation of Verney et al. [61] model for the temporal evolution of mean floc size. Such validation should be expanded for different data sets and for different floc size statistics in order to better constrain the empirical model parameters.

The sensitivity of the modeled mean floc diameter to the prescribed fractal dimension has also been discussed in Verney et al. [61]. More recently, Penaloza-Giraldo et al. [124] report that the temporal evolution of floc sizes (timescale to reach the equilibrium floc sizes) are sensitive to the fractal dimension nf. As shown in Eq. (30), when nf is smaller, the fragmentation probability function becomes larger, and hence it takes a shorter time to reach the equilibrium floc size. Even when α/β and nf are calibrated to match the measured mean equilibrium floc size and flocculation timescale, it may not be straightforward to match the model results with measured data for the entire floc size distribution. Specifically, since the choice of breakup distribution function Πki is not constrained, the calibrated α/β and nf may also depend on the Πki-function. In summary, the sensitivity study of Verney et al. [61] and preliminary findings reported by Penaloza-Giraldo et al. [124] imply that the following flocculation physics controlling the empirical parameters in the PBE warrant future studies:

3.2.1 Parameterizing the floc structure with a fractal dimension

To estimate floc mass and floc density in a given floc size, further assumptions on floc structures is needed. The state-of-the-art model for floc structure is to assume a fractal structure, which renders Eqs. (21) and (22) useful. In the PBE models for floc dynamics, fractal dimension directly affects the breakup term via the fragmentation probability function (see Eq. (30)). While most of the existing flocculation models assume a constant fractal dimension, examining the field and laboratory data by relating measured floc settling velocity and floc size (see Eqs. (22) and (23)) suggests that the fractal dimension may not be a constant, especially when considering the PBE equations are very sensitive to the prescribed fractal dimension value. Recent grain-resolving simulations (see Section 2.2.2 and the studies of [85, 86] referenced therein) also confirms that the fractal dimension depends on floc size. Researchers have proposed to model fractal dimension as a function of floc size [65, 71], however, whether it significantly improves the modeled flocculation processes remain to be proven.

3.2.2 More complete descriptions of collisional efficiency and fragmentation rate

As discussed by Hill and Nowell [125], the collision efficiency is practically treated as an empirical tuning parameter that parameterizes three main processes: encounter, contact and sticking. Encounter and contact are physical processes while sticking is associated with chemical-biological processes. From a physical perspective, only sticking efficiency is solely an empirical parameter. Maggi et al. [71] used a more complex collisional efficiency formulation that depends on the size and porosity of two colliding flocs. Through detailed laboratory experiments, Soos et al. [126] proposed a collisional efficiency formulation αij that depends on the size of the two colliding flocs and the turbulent shear rate. More systematic studies on the impact of collisional efficiency on the modeled flocculation in the PBE-based formulation are required.

While the existing studies mostly treat the fragmentation rate to be a model constant, physically the fragmentation rate βi further depends on the floc breakage force Fy and the dynamic viscosity of the fluid [71]:

βi=EμFy1/2E31

with E an empirical constant. It is clear that the only way to justify that the fragmentation rate is a constant, i.e., βi=β, is to assume that the floc breakage force Fy is independent of floc size, which theoretically is consistent with the fractal theory [69]. Indeed, in most cohesive sediment transport literature (e.g., [61, 71, 115]), Fy is assumed to be Fy=O1010N following Winterwerp [62]. Since assuming floc structure follows the fractal theory (independent of the size of the aggregates) is an approximation, Fy may need to be considered as a variable in practice. Jarvis et al. [127] presented a review of many laboratory measurements of floc breakage force per unit area (it was called floc strength in their paper). Due to a wide range of cohesive sediment samples tested and different measurement techniques used, the quantitative results are not conclusive. However, it is clear that Fy/Df2 decreases as floc size increases. This conclusion does not exclude the possibility that Fy is a variable and it may be insensitive, or has a certain degree of dependence on floc size. More quantitative investigations on floc breakup force for different mineral composition and turbulence intensity are needed.

3.2.3 Improved physical understanding of breakup distribution functions

In the coastal sediment literature, detailed studies on floc breakup, particularly from the observational perspective, are rare. In the water quality literature, Jarvis et al. [127] provide some insights into the breakup distribution function. First, they discern the fragmentation mechanism, similar to the binary breakup, as the most likely scenario to occur when flocs are subjected to tensile stress acting across the floc. Secondly, the erosion mechanism in floc breakup is likely due to shear stress acting tangentially to the floc surface. Based on this argument, researchers hypothesize that the floc breakage types may depend on the ratio of floc size to the Kolmogorov length scale (smallest turbulent eddy size). However, the results are not conclusive and more comprehensive studies on how turbulent eddies interact with flocs and causing floc breakage are warranted.

Advertisement

4. Conclusions

We have presented an overview covering different types of floc analyses based on experimental measurements and grain-resolved simulations. These tools are currently emerging and show a very promising perspective to generate the data needed to account for unresolved cohesive sediment dynamics in continuum models with high fidelity. More work will be needed in the future to cover the different aspects laid out in this chapter. Those are in particular, the effects of biofilms, the settling velocity of different types of flocs, as well as the aggregation and break-up efficiencies governing the exchange between different classes of PBE-type models. The knowledge to be gained can lead to a new generation of continuum models that enable simulations with predictive power for entire estuaries, which will bring inestimable advantages for these attractive settlement areas, both in economic terms and in terms of an increased quality of life.

Advertisement

Acknowledgments

BV gratefully acknowledges the support through the German Research Foundation (DFG) grant VO2413/2-1. KZ is supported by the National Natural Science Foundation of China through the Basic Science Center Program for Ordered Energy Conversion (51888103). LY, AJM, TJH and EM gratefully acknowledge supports by US National Science Foundation through a collaborative research between the University of Delaware (OCE-1924532) and the University of California Santa Barbara (OCE-1924655). AJM’s contribution towards this research was partly supported by the US National Science Foundation under grants OCE-1736668 and OCE-1924532, TKI-MUSA project 11204950-000-ZKS-0002, and HR Wallingford company research FineScale project (ACK3013_62).

Advertisement

Conflict of interest

The authors declare no conflict of interest.

Advertisement

Nomenclature

αij=αcollisional efficiency
βi=βfragmentation rate
Γp,ifluid-particle interface of particle i
δi,jKronecker delta
ζ0characteristic distance
ζngap size in between particles
ζminlimiting gap size in between particles
λcohesive force range
μdynamic fluid viscosity
νkinematic fluid viscosity
νtturbulent viscosity
Πkibreakup distribution function
ρffluid density
ρkffloc density of class k
ρp,idensity of particle i
ρsdensity ratio
ρsprimary particle density
ρwdensity of water
τturbulent shear stress
τhydrodynamic stress tensor
τsnondimensional reference time
ϕparticle volume fraction
ωp,iAngular velocity vector of particle with index i
AijShear-driven binary collision probability function
AhHamaker constant
a1model constant
a2model constant
bagglomeration rate
Bifragmentation probability function
CSPM concentration
cempirical constant
cksediment mass concentration of class k
Cocohesive number
Dfloc size
D50median grain size
D¯fmean floc size
Df,kfloc diameter of class k
Dpprimary particle diameter
Dp,idiameter of particle i
Eempirical constant
Fc,iparticle-interaction force vector of particle with index i
Fcoh,ijcohesive force vector of particle i interacting with particle j
Fcon,ijnormal contact force vector of particle i interacting with particle j
Fh,ihydrodynamic force vector of particle with index i
Fg,igravitational force vector of particle with index i
fIBMartificial volume forced introduced by the IBM i
Flub,ijlubrication force vector of particle i interacting with particle j
Fyfloc breakage force
Gturbulence parameter / shear rate
G1,kmass gain due to aggregation of flocs
Gagain of flocs in a size class due to aggregation
Lbsgain of flocs in a size class due to shear driven break-up
ggravitational acceleration
g′specific gravity
hcothickness of cohesive shell
Ip,imoment of inertia of particle i
kempirical constant
kindex of class for sediment mass concentration
kA′agglomeration constant
kB′breakup constant
L0characteristic length
L1,kmass loss due to breakup of flocs
Laloss of flocs in a size class due to aggregation
Lbsloss of flocs in a size class due to shear driven break-up
mempirical constant
mkmass of an individual floc in class k
Nnumber of classes for sediment mass concentration
nffractal dimension
Nftotal number of flocs in the flow
nknumber concentration of class k
Nptotal number of particles in the flow
N¯p,localaverage number of primary particles per floc
m50mass of the median grain size
mp,imass of particle i
mkmass of a floc
rradial position vector
ReReynolds number
StStokes number
ttime
Tc,iTorque vector of particle with index i due to particle-interaction
Tcon,ijcollision torque vector of particle i interacting with particle j
Th,ihydrodynamic torque vector of particle with index i
Tlub,ijlubrication torque vector of particle i interacting with particle j
U0characteristic velocity
uf,i=ufluid velocity vector
up,itranslational velocity vector of particle with index i
usbuoyancy velocity
vsStokes settling velocity
Wssettling velocity of microflocs
ws,ksettling velocity of of class k
Ws,macrosettling velocity of macroflocs
Advertisement

Abbreviations

ADVacoustic doppler velocimeter
COAWSTCoupled Ocean-Atmosphere-Wave-Sediment Transport Modeling System
DNSdirect numerical simulation
EPSextracellular polymeric substance
ETMestuarine turbidity maximum
FVCOMfinite volume community ocean model
IBMimmersed boundary method
INSSEVIN-Situ SEttling Velocity instrument
LabSFLOClaboratory spectral flocculation charactersitics
MFSVmanning floc settling velocity
MSFmass settling flux
OBSoptical backscatter
PBEpopulation balance equation
PICSparticle imaging camera system
POMPrinceton ocean models
RANSReynolds-averaged Navier-Stokes
ROMSregional ocean modeling systems
SBsize band
SPMsuspended particulate matter
TEPtransparent exopolymer particles

References

  1. 1. Nimmo Smith WAM, Katz J, Osborn TR. On the structure of turbulence in the bottom boundary layer of the coastal ocean. Journal of Physical Oceanography. 2005;35(1):72-93
  2. 2. Verney R, Lafite R, Brun Cottan JC. Flocculation potential of natural estuarine particles: The importance of environmental factors and of the spatial and seasonal variability of suspended particulate matter. Estuaries and Coasts. 2009;32:678-693
  3. 3. Geyer WR, Lavery AC, Scully ME, Trowbridge JH. Mixing by shear instability at high Reynolds number. Geophysical Research Letters. 2010;37:L22607
  4. 4. Foster DL, Beach RA, Holman RA. Turbulence observations of the nearshore wave bottom boundary layer. Journal of Geophysical Research, Oceans. 2006;111(C4):C04011
  5. 5. Ozdemir CE, Hsu TJ, Balachandar S. Direct numerical simulations of transition and turbulence in Stokes boundary layer. Physics of Fluids. 2014;26:045108
  6. 6. Thomson J, Schwendeman MS, Zippel SF, Moghimi S, Gemmrich J, Rogers WE. Wave-breaking turbulence in the ocean surface layer. Journal of Physical Oceanography. 2016;46(6):1857-1870
  7. 7. Lee JH, Monty JP, Elsnab J, Toffoli A, Babanin AV, Alberello A. Estimation of kinetic energy dissipation from breaking waves in the wave crest region. Journal of Physical Oceanography. 2017;47(5):1145-1150
  8. 8. Vowinckel B. Incorporating grainscale processes in macroscopic sediment transport models. Acta Mechanica. 2021;232:2023–2050. DOI: 10.1007/s00707-021-02951-4
  9. 9. Kirby R. Suspended fine cohesive sediment in the Severn estuary and inner Bristol Channel, UK UKAEA Atomic Energy Research Establishment, Harwell. Energy 1986. p. 243
  10. 10. van Leussen W. Fine sediment transport under tidal action. Geo-marine Letters. 1991;11(3):119-126
  11. 11. Manning AJ, Bass SJ. Variability in cohesive sediment settling fluxes: Observations under different estuarine tidal conditions. Marine Geology. 2006;235(1–4):177-192
  12. 12. Dyer KR. Sediment processes in estuaries: Future research requirements. Journal of Geophysical Research. 1989;94(C10):327-332
  13. 13. Eisma D. Flocculation and de-flocculation of suspended matter in estuaries. Netherlands Journal of Sea Research. 1986;20(2–3):183-199
  14. 14. Eisma D, Dyer K, Van Leussen W. The in-situ determination of the settling velocities of suspended fine-grained sediment—A review. In: Burt N, Parker R, Watts J, editors. Cohesive Sediments—Proceedings of INTERCOH Conference. Wallingford, England: John Wiley and Sons; 1997. pp. 17-44
  15. 15. Law D, Bale A, Jones S. Adaptation of focused beam reflectance measurement to in-situ particle sizing in estuaries and coastal waters. Marine Geology. 1997;140(1–2):47-59
  16. 16. Agrawal Y, Pottsmith H. Laser diffraction particle sizing in STRESS. Continental Shelf Research. 1994;14(10–11):1101-1121
  17. 17. Graham GW, Nimmo Smith WAM. The application of holography to the analysis of size and settling velocity of suspended cohesive sediments. Limnology and Oceanography: Methods. 2010;8(1):1-15
  18. 18. Benson T, French J. InSiPID: A new low-cost instrument for in situ particle size measurements in estuarine and coastal waters. Journal of Sea Research. 2007;58(3):167-188
  19. 19. van Leussen W. Estuarine Macroflocs and their Role in Fine-Grained Sediment Transport. The Netherlands: University of Utrecht; 1994
  20. 20. Dearnaley M. Direct measurements of settling velocities in the Owen tube: A comparison with gravimetric analysis. Journal of Sea Research. 1996;36(1–2):41-47
  21. 21. Defossez J. Dynamique des macroflocs au cours de cycles tidaux, Mise au point d’un système d’observation. Rouen, France: VIL, Video In Lab. Mémoire de DEA, Université de Rouen; 1996
  22. 22. Fennessy M, Dyer K, Huntley D. INSSEV: An instrument to measure the size and settling velocity of flocs in situ. Marine Geology. 1994;117(1–4):107-117
  23. 23. Manning AJ, Dyer KR. The use of optics for the in situ determination of flocculated mud characteristics. Journal of Optics A: Pure and Applied Optics. 2002;4(4):S71
  24. 24. Manning AJ, Baugh JV, Spearman JR, Pidduck EL, Whitehouse RJ. The settling dynamics of flocculating mud-sand mixtures: Part 1—Empirical algorithm development. Ocean Dynamics. 2011;61(2–3):311-350
  25. 25. Manning A. LabSFLOC-2—The second generation of the laboratory system to determine spectral characteristics of flocculating cohesive and mixed sediments. In: HR Wallingford Technical Report (TR 156). Wallingford, UK: HR Wallingford; 2015
  26. 26. Manning A. LabSFLOC—A laboratory system to determine the spectral characteristics of flocculating cohesive sediments. In: HR Wallingford Technical Report (TR 156). Wallingford, UK: HR Wallingford; 2006
  27. 27. Ye L, Manning AJ, Hsu TJ, Morey S, Chassignet EP, Ippolito TA. Novel application of laboratory instrumentation characterizes mass settling dynamics of oil-mineral aggregates (OMAs) and oil-mineral-microbial interactions. Marine Technology Society Journal. 2018;52(6):87-90
  28. 28. Manning A, Schoellhamer DH. Factors controlling floc settling velocity along a longitudinal estuarine transect. Marine Geology. 2013;345:266-280
  29. 29. Smith SJ, Friedrichs CT. Size and settling velocities of cohesive flocs and suspended sediment aggregates in a trailing suction hopper dredge plume. Continental Shelf Research. 2011;31(10):S50-S63
  30. 30. ten Brinke WB. Settling velocities of mud aggregates in the Oosterschelde tidal basin (the Netherlands), determined by a submersible video system. Estuarine, Coastal and Shelf Science. 1994;39(6):549-564
  31. 31. Fennessy M, Dyer K. Floc population characteristics measured with INSSEV during the Elbe estuary intercalibration experiment. Journal of Sea Research. 1996;36(1–2):55–62
  32. 32. Manning AJ. Study of the effect of turbulence on the properties of flocculated mud. PhD thesis. UK: Institute of Marine Sciences, Faculty of Science, University of Plymouth; 2001
  33. 33. Stokes GG. On the effect of the internal friction of fluids on the motion of pendulums. Transactions of the Cambridge Philosophical Society. 1851;9:8-106
  34. 34. Manning AJ, Whitehouse RJS, Uncles RJ. Suspended particulate matter: the measurements of flocs. In: Uncles RJ. Mitchell S, editors. ECSA practical handbooks on survey and analysis methods: Estuarine and coastal hydrography and sedimentology, Chapter 8. Cambridge University Press; 2017. pp 211–260. DOI: 10.1017/9781139644426. ISBN 978-1-107-04098-4
  35. 35. Dyer K, Cornelisse J, Dearnaley M, Fennessy M, Jones S, Kappenberg J, et al. A comparison of in situ techniques for estuarine floc settling velocity measurements. Journal of Sea Research. 1996;36(1–2):15-29
  36. 36. Krone RB. A Study of Rheological Properties of Estuarial Sediments. Berkeley: Hydraulic Engineering Laboratory and Sanitary Engineering Research Laboratory, University of California; 1963
  37. 37. Kolmogorov AN. The local structure of turbulence in incompressible viscous fluid for very large Reynolds numbers. Proceedings of the Royal Society of London Series A: Mathematical and Physical Sciences. 1890;1941(434):9–13
  38. 38. Kolmogorov AN. Dissipation of energy in the locally isotropic turbulence. Doklady Akademii Nauk SSSR. 1941;32:16-18
  39. 39. Mehta AJ, Lott JW. Sorting of fine sediment during deposition. In: Kraus NC, editor. Coastal Sediments ’87, vol. I. New York: American Society of Civil Engineers; 1987. pp. 348–362
  40. 40. Glasgow LA, Luecke RH. Mechanisms of deaggregation for clay-polymer flocs in turbulent systems. Industrial & Engineering Chemistry Fundamentals. 1980;19(2):148-156
  41. 41. McCave IN. Size spectra and aggregation of suspended particles in the deep ocean. Deep Sea Research. 1984;31:329-352
  42. 42. Alldredge AL, Gotschalk C. In situ settling behavior of marine snow 1. Limnology and Oceanography. 1988;33(3):339-351
  43. 43. Manning A, Friend P, Prowse N, Amos C. Preliminary findings from a study of Medway Estuary (UK) natural mud floc properties using a laboratory mini-flume and the LabSFLOC system. Continental Shelf Research, BIOFLOW SI. 2007:1080-1095
  44. 44. Overbeek JTG, Jonker G. Colloid Science Edited by HR Kruyt: Contributors-GH Jonker, J. Th. G. Overbeek. Amsterdam, The Netherlands: Elsevier; 1952
  45. 45. Mikeš D, Manning AJ. Assessment of flocculation kinetics of cohesive sediments from the seine and Gironde estuaries, France, through laboratory and field studies. Journal of Waterway, Port, Coastal, and Ocean Engineering. 2010;136(6):306-318
  46. 46. Pouët M.-F. La Clarification Coagulation—Flocculation Traitement de l’eau potable cours, EMA, option Environnement et Systèmes Industriels. 1997
  47. 47. Manning AJ, Baugh JV, Spearman JR, Whitehouse RJ. Flocculation settling characteristics of mud: Sand mixtures. Ocean Dynamics. 2010;60(2):237-253
  48. 48. Whitehouse RJS, Soulsby R, Roberts W, Mitchener HJ. Dynamics of Estuarine Muds. London: Thomas Telford Publications; 2000
  49. 49. Delo EA, Ockenden MC. Estuarine muds manual. Report SR 309. Wallingford: HR Wallingford; 1992
  50. 50. Odd NVM, Roger JG. An analysis of the behaviour of fluid mud in estuaries. HR Wallingford Technical Report, SR 84. 1986:94
  51. 51. Soulsby RL. Methods for predicting suspensions of mud. In: HR Wallingford Technical Report (TR 104). Wallingford, UK: HR Wallingford; 2000
  52. 52. Argaman Y, Kaufman WJ. Turbulence and flocculation. Journal of the Sanitary Engineering Division. 1970;96(2):223-241
  53. 53. Lick W, Huang H, Jepsen R. Flocculation of fine-grained sediments due to differential settling. Journal of Geophysical Research: Oceans. 1993;98(C6):10279-10288
  54. 54. Manning A, Dyer K. Mass settling flux of fine sediments in northern European estuaries: Measurements and predictions. Marine Geology. 2007;245(1–4):107-122
  55. 55. Manning AJ. The observed effects of turbulence on estuarine flocculation. Sediment Transport in European Estuarine Environments: Proceedings of the STRAEE Workshop (WINTER 2004). Journal of Coastal Research. 2004;(Special issue 41):90-104
  56. 56. Manning A, Spearman J, Whitehouse R, Pidduck E, Baugh J, Spencer K. Laboratory assessments of the flocculation dynamics of mixed mud: Sand suspensions. Chapter. 2013;6:119-164
  57. 57. Baugh JV, Manning AJ. An assessment of a new settling velocity parameterisation for cohesive sediment transport modeling. Continental Shelf Research. 2007;27(13):1835-1855
  58. 58. Mietta F, Chassagne C, Manning AJ, Winterwerp JC. Influence of shear rate, organic matter content, pH and salinity on mud flocculation. Ocean Dynamics. 2009;59(5):751-763
  59. 59. Soulsby RL, Manning AJ, Spearman J, Whitehouse RJS. Settling velocity and mass settling flux of flocculated estuarine sediments. Marine Geology. 2013;339:1-12
  60. 60. Lee BJ, Toorman E, Molz FJ, Wang J. A two-class population balance equation yielding bimodal flocculation of marine or estuarine sediments. Water Research. 2011;45(5):2131-2145
  61. 61. Verney R, Lafite R, Brun-Cottan JC, Le Hir P. Behaviour of a floc population during a tidal cycle: Laboratory experiments and numerical modeling. Continental Shelf Research. 2011;31:S64-S83
  62. 62. Winterwerp JC. A simple model for turbulence induced flocculation of cohesive sediment. Journal of Hydraulic Research. 1998;36(3):309-326
  63. 63. Winterwerp H. On the dynamics of high-concentrated mud suspensions. Communications on Hydraulic and Geotechnical Engineering. 1999
  64. 64. Winterwerp J, Manning A, Martens C, De Mulder T, Vanlede J. A heuristic formula for turbulence-induced flocculation of cohesive sediment. Estuarine, Coastal and Shelf Science. 2006;68(1–2):195-207
  65. 65. Khelifa A, Hill PS. Models for effective density and settling velocity of flocs. Journal of Hydraulic Research. 2006;44:390-401
  66. 66. Maggi F. Variable fractal dimension: A major control for floc structure and flocculation kinematics of suspended cohesive sediment. Journal of Geophysical Research. 2007;112:C07012
  67. 67. Son M, Hsu TJ. The effect of variable yield strength and variable fractal dimension on flocculation of cohesive sediment. Water Research. 2009;43(14):3582-3592
  68. 68. Engel A, Schartau M. Influence of transparent exopolymer particles (TEP) on sinking velocity of Nitzschia closterium aggregates. Marine Ecology Progress Series. 1999;182:69-76
  69. 69. Kranenburg C. The fractal structure of cohesive sediment aggregates. Estuarine, Coastal and Shelf Science. 1994;39(5):451-460
  70. 70. McAnally WH, Mehta JA. Significance of aggregation of fine sediment particles in their deposition. Estuarine, Coastal and Shelf Science. 2002;54:643-653
  71. 71. Maggi F, Mietta F, Winterwerp JC. Effect of variable fractal dimension on the floc size distribution of suspended cohesive sediment. Journal of Hydrology. 2007;343:43-55
  72. 72. Corning PA. The re-emergence of “emergence”: A venerable concept in search of a theory. Complexity. 2002;7(6):18-30
  73. 73. Chialvo DR. Emergent complex neural dynamics. Nature Physics. 2010;6(10):744-750
  74. 74. Halley JD, Winkler DA. Classification of emergence and its relation to self-organization. Complexity. 2008;13(5):10-15
  75. 75. Cranford SW, De Boer J, Van Blitterswijk C, Buehler MJ. Materiomics: An-omics approach to biomaterials research. Advanced Materials. 2013;25(6):802-824
  76. 76. Rietkerk M, Van de Koppel J. Regular pattern formation in real ecosystems. Trends in Ecology & Evolution. 2008;23(3):169-175
  77. 77. Spencer KL, Wheatland JA, Bushby AJ, Carr SJ, Droppo IG, Manning AJ. A structure–function based approach to floc hierarchy and evidence for the non-fractal nature of natural sediment flocs. Scientific Reports. 2021;11(1):1-10
  78. 78. Spencer KL, Manning AJ, Droppo IG, Leppard GG, Benson T. Dynamic interactions between cohesive sediment tracers and natural mud. Journal of Soils and Sediments. 2010;10(7):1401-1414
  79. 79. Wheatland JA, Bushby AJ, Spencer KL. Quantifying the structure and composition of flocculated suspended particulate matter using focused ion beam nanotomography. Environmental Science & Technology. 2017;51(16):8917-8925
  80. 80. Wheatland JA, Spencer KL, Droppo IG, Carr SJ, Bushby AJ. Development of novel 2D and 3D correlative microscopy to characterise the composition and multiscale structure of suspended sediment aggregates. Continental Shelf Research. 2020;200:104112
  81. 81. Zhang N, Thompson CE, Townend IH, Rankin KE, Paterson DM, Manning AJ. Nondestructive 3D imaging and quantification of hydrated biofilm-sediment aggregates using X-ray microcomputed tomography. Environmental Science & Technology. 2018;52(22):13306-13313
  82. 82. Ye L, Manning AJ, Hsu TJ. Oil-mineral flocculation and settling velocity in saline water. Water Research. 2020;173:115569
  83. 83. Biegert E, Vowinckel B, Meiburg E. A collision model for grain-resolving simulations of flows over dense, mobile, polydisperse granular sediment beds. Journal of Computational Physics. 2017;340:105-127
  84. 84. Biegert E, Vowinckel B, Ouillon R, Meiburg E. High-resolution simulations of turbidity currents. Progress in Earth and Planetary Science. 2017;4(1):33
  85. 85. Zhao K, Vowinckel B, Hsu TJ, Köllner T, Bai B, Meiburg E. An efficient cellular flow model for cohesive particle flocculation in turbulence. Journal of Fluid Mechanics. 2020;889:R3
  86. 86. Zhao K, Pomes F, Vowinckel B, Hsu TJ, Bai B, Meiburg E. Flocculation of suspended cohesive particles in homogeneous isotropic turbulence. Journal of Fluid Mechanics. 2021;921:A17
  87. 87. Cox RG, Brenner H. The slow motion of a sphere through a viscous fluid towards a plane surface—II small gap widths, including inertial effects. Chemical Engineering Science. 1967;22(12):1753-1777
  88. 88. Vowinckel B, Withers J, Luzzatto-Fegiz P, Meiburg E. Settling of cohesive sediment: Particle-resolved simulations. Journal of Fluid Mechanics. 2019;858:5-44
  89. 89. Maxey MR. The motion of small spherical particles in a cellular flow field. Physics of Fluids. 1987;30:1915-1928
  90. 90. Bergougnoux L, Bouchet G, Lopez D, Guazzelli E. The motion of solid spherical particles falling in a cellular flow field at low Stokes number. Physics of Fluids. 2014;26(9):093302
  91. 91. Kuprenas R, Tran D, Strom K. A shear-limited flocculation model for dynamically predicting average floc size. Journal of Geophysical Research-Oceans. 2018;123:6736-6752
  92. 92. Tran D, Kuprenas R, Strom K. How do changes in suspended sediment concentration alone influence the size of mud flocs under steady turbulent shearing? Continental Shelf Research. 2018;158:1-14
  93. 93. Vowinckel B, Biegert E, Luzzatto-Fegiz P, Meiburg E. Consolidation of freshly deposited cohesive and noncohesive sediment: Particle-resolved simulations. Physical Review Fluids. 2019;4(7):074305
  94. 94. Winterwerp JC. On the flocculation and settling velocity of estuarine mud. Continental shelf research. 2002;22(9):1339-1360
  95. 95. Uhlmann M. An immersed boundary method with direct forcing for the simulation of particulate flows. Journal of Computational Physics. 2005;209(2):448-476
  96. 96. Kempe T, Fröhlich J. An improved immersed boundary method with direct forcing for the simulation of particle laden flows. Journal of Computational Physics. 2012;231(9):3663-3684
  97. 97. Winterwerp JC, Kranenburg C. Fine Sediment Dynamics in the Marine Environment. Amsterdam, The Netherlands: Elsevier; 2002
  98. 98. Warner JC, Armstrong B, He R, Zambon JB. Development of a coupled ocean-atmosphere-wave-sediment transport (COAWST) modeling system. Ocean Modelling. 2010;35:230-244
  99. 99. Ge J, Shen F, Guo W, Chen C, Ding P. Estimation of critical shear stress for erosion in the Changjiang Estuary: A synergy research of observation, GOCI sensing and modeling. Journal of Geophysical Research, Oceans. 2015;120:8439-8465
  100. 100. van Maren DS, Winterwerp JC, Vroom J. Fine sediment transport into the hyper-turbid lower Ems River: The role of channel deepening and sediment-induced drag reduction. Ocean Dynamics. 2015;65:589-605
  101. 101. Booij N, Ris RC, Holthuijsen LH. A third generation wave model for coastal regions. Journal of Geophysical Research. 1999;104(C4):7649-7666
  102. 102. Grant WD, Madsen OS. The continental-shelf bottom boundary layer. Annual Review of Fluid Mechanics. 1986;18(1):265-305
  103. 103. Milligan T, Hill P. A laboratory assessment of the relative importance of turbulence, particle composition, and concentration in limiting maximal floc size and settling behaviour. Journal of Sea Research. 1998;39(3-4):227-241
  104. 104. Keyvani SKA. Influence of cycles of high and low turbulent shear on the growth rate and equilibrium size of mud flocs. Marine Geology. 2014;354:1-14
  105. 105. Tran C, Strom K. Suspended clays and silts: Are they independent or dependent fractions when it comes to settling in a turbulent suspension? Continental Shelf Research. 2017;138:81-94
  106. 106. Fugate DC, Friedrichs CT. Controls on suspended aggregate size in partially mixed estuaries. Estuarine, Coastal and Shelf Sciences. 2003;58:389-404
  107. 107. Fettweis M, Francken F, Pison V, Van den Eynde D. Suspended particulate matter dynamics and aggregate sizes in a high turbidity area. Marine Geology. 2006;235:63-74
  108. 108. Manning AJ, Dyer KR. A laboratory examination of floc characteristics with regard to turbulent shearing. Marine Geology. 1999;160(1-2):147-170
  109. 109. Manning AJ, Bass SJ, Dyer KR. Floc properties in the turbidity maximum of a mesotidal estuary during neap and spring tidal conditions. Marine Geology. 2006;235(1-4):193-211
  110. 110. Cartwright GM, Friedrichs CT, Sanford LP. In situ characterization of estuarine suspended sediment in the presence of muddy flocs and pellets. The Proceedings of the Coastal Sediments. 2011;2011:642-655
  111. 111. Malpezzi MA, Sanford LP, Crump B. Abundance and distribution of transparent exopolymer particles in the estuarine turbidity maximum of Chesapeake Bay. Marine Ecology Progress Series. 2013;486:23-35
  112. 112. Manning AJ, Baugh JV, Spearman JR, Pidduck EL, Whitehouse RJS. Flocculation settlingCharacteristics of mud: Sand mixtures. Ocean Dynamics. 2010;60:237-253
  113. 113. Cuthbertson AJS, Samsami F, Dong P. Model studies for flocculation of sand-clay mixtures. Coastal Engineering. 2018;132:13-32
  114. 114. Xu F, Wang DP, Riemer N. Modeling flocculation processes of fine- grained particles using a size-resolved method: Comparison with published laboratory experiments. Continental Shelf Research. 2008;28:2668-2677
  115. 115. Xu F, Wang DP, Riemer N. An idealized model study of flocculation on sediment trapping in an estuarine turbidity maximum. Continental Shelf Research. 2010;30:1314-1323
  116. 116. Sherwood CR, Aretxabaleta AL, Harris CK, Rinehimer JP, Verney R, Ferré B. Cohesive and mixed sediment in the Regional Ocean Modeling system (ROMS v3.6) implemented in the coupled ocean–atmosphere–wave–sediment transport Modeling system (COAWST r1234). Geoscientific Model Development. 2018;11:1849-1871
  117. 117. Shen X, Lee BJ, Fettweis M, Toorman EA. A tri-modal flocculation model coupled with TELEMAC for estuarine muds both in the laboratory and in the field. Water Research. 2018;145:473-486
  118. 118. Liu J, Liang JH, Xu K, Chen Q, Chen Q, Ozdemir CE. Modeling sediment flocculation in Langmuir turbulence. Journal of Geophysical Research: Oceans. 2019;124:7883-7907
  119. 119. Shen X, Maa JPY. Modeling floc size distribution of suspended cohesive sediments using quadrature method of moments. Marine Geology. 2015;359:106-119
  120. 120. Maerz J, Verney R, Wirtz K, Feudel U. Modeling flocculation processes: Intercomparison of a size class-based model and a distribution-based model. Continental Shelf Research. 2011;31:S84-S93
  121. 121. Son M, Hsu TJ. The effects of flocculation and bed erodibility on modeling cohesive sediment resuspension. Journal of Geophysical Research. 2011;116:C03021
  122. 122. Shen X, Maa JPY. Numerical simulations of particle size distributions: Comparison with analytical solutions and kaolinite flocculation experiments. Marine Geology. 2016;379:84-99
  123. 123. Warner JC, Sherwood CR, Signell RP, Harris CK, Arango HG. Development of a three-dimensional, regional, coupled wave, current, and sediment-transport model. Computational Geosciences. 2008;34(10):1243-1260
  124. 124. Penaloza-Giraldo J, Yue L, Ye L, Hsu TJ, Manning AJ, Meiburg E H, Vowinckel B. The Effect of Floc Strength in a Size Class-Based Flocculation Model. In AGU Fall Meeting Abstracts. San Francisco, USA; Dec 2020;2020:EP001-0020
  125. 125. Hill PS, Nowell ARM. The potential role of large, fast-sinking particles in clearing nepheloid layers. Philosophical Transactions of the Royal Society of London A. 1990;331:103-117
  126. 126. Soos M, Wang L, Fox RO, Sefcik J, Morbidelli M. Population balance modeling of aggregation and breakage in turbulent Taylor–Couette flow. Journal of Colloid and Interface Science. 2007;307:433-446
  127. 127. Jarvis P, Jefferson B, Gregory J, Parsons SA. A review of floc strength and breakage. Water Research. 2005;39:3121-3137

Written By

Bernhard Vowinckel, Kunpeng Zhao, Leiping Ye, Andrew J. Manning, Tian-Jian Hsu, Eckart Meiburg and Bofeng Bai

Reviewed: 01 March 2022 Published: 03 May 2022