Skip to main content
Log in

Characterization of bedload intermittency near the threshold of motion using a Lagrangian sediment transport model

  • Original Article
  • Published:
Environmental Fluid Mechanics Aims and scope Submit manuscript

Abstract

At the smallest scales of sediment transport in rivers, the coherent structures of the turbulent boundary layer constitute the fundamental mechanisms of bedload transport, locally increasing the instantaneous hydrodynamic forces acting on sediment particles, and mobilizing them downstream. Near the critical threshold for initiating sediment motion, the interactions of the particles with these unsteady coherent structures and with other sediment grains, produce localized transport events with brief episodes of collective motion occurring due to the near-bed velocity fluctuations. Simulations of these flows pose a significant challenge for numerical models aimed at capturing the physical processes and complex non-linear interactions that generate highly intermittent and self-similar bedload transport fluxes. In this investigation we carry out direct numerical simulations of the flow in a rectangular flat-bed channel, at a Reynolds number equal to Re = 3632, coupled with the discrete element method to simulate the dynamics of spherical particles near the bed. We perform two-way coupled Lagrangian simulations of 48,510 sediment particles, with 4851 fixed particles to account for bed roughness. Our simulations consider a total of eight different values of the non-dimensional Shields parameter to study the evolution of transport statistics. From the trajectory and velocity of each sediment particle, we compute the changes in the probability distribution functions of velocities, bed activity, and jump lengths as the Shields number increases. For the lower shear stresses, the intermittency of the global bedload transport flux is described by computing the singularity or multifr actal spectrum of transport, which also characterizes the widespread range of transport event magnitudes. These findings can help to identify the mechanisms of sediment transport at the particle scale. The statistical analysis can also be used as an ingredient to develop larger, upscaled models for predicting mean transport rates, considering the variability of entrainment and deposition that characterizes the transport near the threshold of motion.

This is a preview of subscription content, log in via an institution to check access.

Access this article

Price excludes VAT (USA)
Tax calculation will be finalised during checkout.

Instant access to the full article PDF.

Fig. 1
Fig. 2
Fig. 3
Fig. 4
Fig. 5
Fig. 6
Fig. 7
Fig. 8
Fig. 9
Fig. 10
Fig. 11

Similar content being viewed by others

Abbreviations

C :

Volumetric concentration of particles

\(C_D\) :

Drag coefficient

d :

Particles diameter

\(D_0\) :

Fractal, Hausdorff or box-counting dimension of the bedload flux

\(d^+\) :

Non-dimensional diameter

e :

Coefficient of restitution

\(F^{col}\) :

Force due to inter-particles collisions

\(F_i\) :

Feedback force from the particles to the flow

Fr :

Froude number

\(f_i\) :

External pressure gradient applied to the flow

\(f_{i,\eta }^{p}\) :

Hydrodynamic drag force felt by a particle

\(f(\alpha )\) :

Singularity spectrum

g :

Gravity acceleration

h :

Length scale of a domain

k :

Shape parameter of the Gamma distribution function

\(M(q,r\Delta t)\) :

q-th Order statistical moment of \(\mu _i\)

\(N(r\Delta t)\) :

Number of time windows of size \(r\Delta t\) contained in the total time when transport occurs

P :

Pressure field of the flow

Re :

Reynolds number based on bulk velocity

\(Re_p\) :

Particle Reynolds number based on the particle velocity relative to the flow

\(Re_p^*\) :

Particle Reynolds number based on the friction velocity

\(Re_\tau \) :

Friction Reynolds number

r :

Integer for computing different sampling time scale of sediment transport

\(r^+_{inactive}\) :

Dimensionless radius of the inactive particles

S(t):

Cumulative amount of particles that have crossed the control plane until time t

\(St_k\) :

Stokes number

s(t):

Number of particles crossing the control plane in a given time step \(\Delta t\)

\(T_{max}\) :

Total time of simulation

\(T(r\Delta t)\) :

Total width of periods with no transport that are larger than \(r\Delta t\)

t :

Time

U :

Bulk velocity of the flow

\(u_{f_i}\) :

Velocity of the fluid at the particle location (\(i=1,2,3\))

\(u_j\) :

Instantaneous velocity components of the flow (\(j=1,2,3\))

\(u_p\) :

Particles velocity in the streamwise direction

\(u^+\) :

Dimensionless bulk velocity

\(u^*\) :

Friction velocity

\(v_{cr}\) :

Arbitrary critical velocity which delineate active motions

\(v_p\) :

Particles velocity in the spanwise direction

\(v_{p_i}\) :

Instantaneous particle velocity component (\(i=1,2,3\))

\(w_p\) :

Particles velocity in the vertical direction

\(w_{\eta }^{j}\) :

Linear geometric weight for the projection of the feedback forces from particles to flow

\(x_i\) :

Cartesian coordinates (\(i=1,2,3\)). Also written as xyz

\(x_L\) :

Streamwise dimension of the channel

Y :

Young’s modulus

\(y_L\) :

Spanwise dimension of the channel

\(z_L\) :

Vertical dimension of the channel

\(z^+\) :

Dimensionless vertical coordinate

\(\alpha \) :

Hölder exponent

\(\beta \) :

Scale parameter of the Gamma distribution function

\(\Delta t\) :

Time step used by the flow solver

\(\Delta t_p\) :

Time step used by the particles solver

\(\Delta V_{\eta }\) :

Volume of a computational cell

\(\delta \) :

Boundary layer thickness

\(\zeta \) :

Scaling exponent function

\(\theta \) :

Shields parameter

\(\vartheta \) :

Poisson’s ratio

\(\kappa \) :

von Kármán constant

\(\lambda \) :

Distance traveled downstream by the particles

\(\mu _f\) :

Coefficient of friction

\(\mu _i\) :

Mass fraction of sediment that cross the reference plane between two successive plateaus

\(\nu \) :

Kinematic viscosity of the fluid

\(\rho _f\) :

Density of the fluid

\(\rho _p\) :

Density of the particles

\(\tau _0\) :

Bed shear stress

\(\mathcal {X}\) :

Random variable distributing Gamma

\(\forall _p\) :

Volume of a particle

References

  1. Adrian RJ (2007) Hairpin vortex organization in wall turbulence. Phys Fluids 19(4):041301

    Article  Google Scholar 

  2. Allen JRL (1971) Bed forms due to mass transfer in turbulent flows: a kaleidoscope of phenomena. J Fluid Mech 49(1):49–63

    Article  Google Scholar 

  3. Alletto M, Breuer M (2012) One-way, two-way and four-way coupled LES predictions of a particle-laden turbulent flow at high mass loading downstream of a confined bluff body. Int J Multiph Flow 45:70–90

    Article  Google Scholar 

  4. Ancey C, Böhm T, Jodeau M, Frey P (2006) Statistical description of sediment transport experiments. Phys Rev E 74(1):011302

    Article  Google Scholar 

  5. Armenio V, Fiorotto V (2001) The importance of the forces acting on particles in turbulent flows. Phys Fluids 13(8):2437–2440

    Article  Google Scholar 

  6. Bak P (1986) The Devil’s staircase. Phys Today 39:38–45

    Article  Google Scholar 

  7. Balachandar S (2013) Particle-laden flows. In: Fernando HJS (ed) Handbook of environmental fluid dynamics, vol 1. CRC Press/Taylor & Francis Group, New York. LLC. ISBN: 978-1-4398-1669-1

  8. Ballio F, Campagnol J, Nikora V, Radice A (2013) Diffusive properties of bed load moving sediments at short time scales. In: Proceedings of 2013 IAHR world congress, Chengdu, China (CD)

  9. Bialik RJ (2013) Numerical study of near-bed turbulence structures influence on the initiation of saltating grains movement. J Hydrol Hydromech 61(3):202–207

    Article  Google Scholar 

  10. Bialik RJ, Nikora V, Karpiński M, Rowiński P (2015) Diffusion of bedload particles in open-channel flows: distribution of travel times and second-order statistics of particle trajectories. Environ Fluid Mech 15(6):1281–1292

    Article  Google Scholar 

  11. Campagnol J, Radice A, Ballio F, Nikora V (2015) Particle motion and diffusion at weak bed load: accounting for unsteadiness effects of entrainment and disentrainment. J Hydraul Res 53(5):633–648

    Article  Google Scholar 

  12. Clift R, Grace J, Weber M (1978) Bubbles, drops, and particles. Academic Press, New York

    Google Scholar 

  13. Coles DE (1956) The law of the wake in the turbulent boundary layer. J Fluid Mech 1:191–226

    Article  Google Scholar 

  14. Coles, DE, Hirst EA (1968) Computation of turbulent boundary layers. In: Proceedings AFOSR-IFP Stanford Conference, Stanford University, Stanford, California, vol 2

  15. Crowe CT, Troutt TR, Chung JN (1996) Numerical models for two-phase turbulent flows. Annu Rev Fluid Mech 28(1):11–43

    Article  Google Scholar 

  16. Chand R, Khaskheli MA, Qadir A, Ge B, Shi Q (2012) Discrete particle simulation of radial segregation in horizontally rotating drum: effects of drum-length and non-rotating end-plates. Phys A 391(20):4590–4596

    Article  Google Scholar 

  17. Chang Y, Scotti A (2003) Entrainment and suspension of sediments into a turbulent flow over ripples. J Turbul 4:019

    Article  Google Scholar 

  18. Dallali M, Armenio V (2015) Large eddy simulation of two-way coupling sediment transport. Adv Water Resour 81:33–44

    Article  Google Scholar 

  19. de Anna P, Le Borgne T, Dentz M, Tartakovsky AM, Bolster D, Davy P (2013) Flow intermittency, dispersion, and correlated continuous time random walks in porous media. Phys Rev Lett 110(18):184502

    Article  Google Scholar 

  20. Drake TG, Shreve RL, Dietrich WE, Whiting PJ, Leopold LB (1988) Bedload transport of fine gravel observed by motion-picture photography. J Fluid Mech 192:193–217

    Article  Google Scholar 

  21. Einstein HA (1937) The bed load transport as probability problem. Ph.D Thesis. Mitt. der Versuchsanst. Wasserbau Eidg. Technical Hochsch, Zürich

  22. Escauriaza C, Sotiropoulos F (2009) Trapping and sedimentation of inertial particles in three-dimensional flows in a cylindrical container with exactly counter-rotating lids. J Fluid Mech 641:169–193

    Article  Google Scholar 

  23. Escauriaza C, Sotiropoulos F (2011) Lagrangian model of bed-load transport in turbulent junction flows. J Fluid Mech 666:36–76

    Article  Google Scholar 

  24. Frey P, Church M (2009) How river beds move. Science 325(5947):1509–1510

    Article  Google Scholar 

  25. Frisch U, Parisi G (1985) On the singularity structure of fully developed turbulence. In: Gil M, Benzi R, Parisi G (eds) Turbulence and predictability in geophysical fluid dynamics. North Holland, New York, pp 84–88

    Google Scholar 

  26. Furbish DJ, Schmeeckle MW (2013) A probabilistic derivation of the exponential-like distribution of bed load particle velocities. Water Resour Res 49(3):1537–1551

    Article  Google Scholar 

  27. Grass AJ (1971) Structural features of turbulent flows over smooth and rough boundaries. J Fluid Mech 50(02):233–255

    Article  Google Scholar 

  28. Harte D (2001) Multifractals: theory and applications. Chapman and Hall/CRC Press, Boca Raton

    Book  Google Scholar 

  29. Heathershaw AD, Thorne PD (1985) Sea-bed noises reveal role of turbulent bursting phenomenon in sediment transport by tidal currents. Nature 316(6026):339–342

    Article  Google Scholar 

  30. Keshavarzi A, Ball J, Nabav H (2012) Frequency pattern of turbulent flow and sediment entrainment over ripples using image processing. Hydrol Earth Syst Sci 16(1):147–156

    Article  Google Scholar 

  31. Keylock CJ, Nishimura K, Peinke J (2012) A classification scheme for turbulence based on the velocity-intermittency structure with an application to near-wall flow and with implications for bed load transport. J Geophys Res 117(F1):F01037

    Article  Google Scholar 

  32. Kézdi Á (1974) Handbook of soil mechanics, vol One. Elsevier, Amsterdam

    Google Scholar 

  33. Le Moigne A, Qin N (2006) LES with numerical dissipation for aircraft wake vortices. In: 44th AIAA aerospace sciences meeting and exhibit, aerospace sciences meetings

  34. Lee H, Ha MY, Balachandar S (2012) Work-based criterion for particle motion and implication for turbulent bed-load transport. Phys Fluids 24(11):116604

    Article  Google Scholar 

  35. Link O, González C, Maldonado M, Escauriaza C (2012) Coherent structure dynamics and sediment particle motion around a cylindrical pier in developing scour holes. Acta Geophys 60(6):1689–1719

    Article  Google Scholar 

  36. Lu SS, Willmarth WW (1973) Measurements of the structures of the Reynolds stress in a turbulent boundary layer. J Fluid Mech 60:481–571

    Article  Google Scholar 

  37. Marion A, Tregnaghi M (2013) A new theoretical framework to model incipient motion of sediment grains and implications for the use of modern experimental techniques. Earth and Planetary Sciences, GeoPlanet, pp 85–100

  38. Michaelides EE (2003) Hydrodynamic force and heat/mass transfer from particles, bubbles, and drops. J Fluids Eng 125:209–238

    Article  Google Scholar 

  39. Moreno-Casas PA, Bombardelli FA (2015) Computation of the Basset force: recent advances and environmental flow applications. Environ Fluid Mech 16(1):193–208

    Article  Google Scholar 

  40. Nelson JM, Shreve RL, McLean SR, Drake TG (1995) Role of near-bed turbulence structure in bed load transport and form mechanics. Water Resour Res 31(8):2071–2086

    Article  Google Scholar 

  41. Nikora V, Habersack H, Huber T, McEwan I (2002) On bed particle diffusion in gravel bed flows under weak bed load transport. Water Resour Res 38(6):1081

    Article  Google Scholar 

  42. Niño Y, García M (1996) Experiments on particle-turbulence interactions in the near-wall region of an open channel flow: implications for sediment transport. J Fluid Mech 326:285–319

    Article  Google Scholar 

  43. Richter DH, Sullivan PP (2013) Momentum transfer in a turbulent, particle-laden Couette flow. Phys Fluids 25:053304

    Article  Google Scholar 

  44. Roseberry JC, Schmeeckle MW, Furbish DJ (2012) A probabilistic description of the bed load sediment flux: 2. Particle activity and motions. J Geophys Res 117(F3):F03032

    Article  Google Scholar 

  45. Schmeeckle MW (2014) Numerical simulation of turbulence and sediment transport of medium sand. J Geophys Res 119:1240–1262

    Article  Google Scholar 

  46. Singh A, Fienberg K, Jerolmack DJ, Marr J, Foufoula-Georgiou E (2009) Experimental evidence for statistical scaling and intermittency in sediment transport rates. J Geophys Res 114(F1):F01025

    Article  Google Scholar 

  47. Soldati A, Marchioli C (2012) Sediment transport in steady turbulent boundary layers: potentials, limitations, and perspectives for Lagrangian tracking in DNS and LES. Adv Water Resour 48:18–30

    Article  Google Scholar 

  48. Spalding DB (1961) A single formula for the law of the wall. J Apply Mech 28:455–458

    Article  Google Scholar 

  49. Sturm TW (2001) Open channel hydraulics (International Edition ed.). McGraw Hill, New York

    Google Scholar 

  50. van der Geld CWM (1997) Measurement and prediction of solid sphere trajectories in accelerated gas flow. Int J Multiph Flow 23(2):357–376

  51. Vanoni V (1975) Sedimentation engineering In: ASCE task committee for the preparation of the manual on sedimentation of the sedimentation committee of the hydraulics division

  52. Vreman AW (2015) Turbulence attenuation in particle-laden flow in smooth and rough channels. J Fluid Mech 773:103–136

    Article  Google Scholar 

  53. Vowinckel B, Kempe T, Fröhlich J (2014) Fluid-particle interaction in turbulent open channel flow with fully-resolved mobile beds. Adv Water Resour 72:32–44

    Article  Google Scholar 

  54. White FM (1991) Viscous fluid flow, vol 2. McGraw Hill, New York

    Google Scholar 

  55. Wu FC, Jiang MR (2007) Numerical investigation of the role of turbulent bursting in sediment entrainment. J Hydraul Eng 133:329–334

    Article  Google Scholar 

Download references

Acknowledgments

This work has been supported by Fondecyt Project 1130940, ONR-G NICOP N622909-11-1-7041, and Conicyt/Fondap Grant 15110017. C. González acknowledges the funding from the Ph.D. National Grant Conicyt-21120939. S. Bateman and J. Calantoni were supported under base funding to the Naval Research Laboratory from the Office of Naval Research.

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to Cristián Escauriaza.

Appendices

Appendix 1

The discrete element method (DEM) used for the computation of the particles dynamics is the open source LIGGGHTS, available on http://www.liggghts.com. The equations solved by this algorithm have been previously explained by Chand et al. [16] and Schmeeckle [45], but following their work and for the clarity of the reader, we repeat them here.

For each particle, collisions with other grains are computed when the distance between their center is less or equal than the sum of their radii, i.e. they overlap, or at least, they are in contact. For two particles, the force product of their collision is divided into a normal force (\(F_n\)) and a tangencial force (\(F_t\)):

$$\begin{aligned} F^{col} = F_n + F_t \end{aligned}$$
(15)

The normal force along the line of center between two colliding particles is computed as:

$$\begin{aligned} F_n = K_n \delta _n - \gamma _n v_n \end{aligned}$$
(16)

whereas the tangencial force is calculated as follows:

$$\begin{aligned} F_t = K_t \delta _t - \gamma _t v_t \end{aligned}$$
(17)

where \(K_n\) and \(K_t\) are the stiffness coefficients, \(\delta _n\) and \(\delta _t\) are the overlap distance between two grains, \(\gamma _n\) and \(\gamma _t\) are the viscoelastic damping constants, \(v_n\) and \(v_t\) the relative velocity and the subscripts n and t correspond to the normal and tangencial components. If necessary, the tangential overlap is truncated in order to satisfy the condition \(F_t \le \mu _f F_n\), where \(\mu _f\) is the coefficient of friction. In turn, the previous variables are computed as follows:

$$\begin{aligned} K_n = \frac{4}{3} Y^* \sqrt{R^* \delta _n} \end{aligned}$$
(18)
$$\begin{aligned} \gamma _n = -2 \sqrt{\frac{5}{6}} B \sqrt{S_n m*} \ge 0 \end{aligned}$$
(19)
$$\begin{aligned} K_t = 8 G^* \sqrt{R^* \delta _n} \end{aligned}$$
(20)
$$\begin{aligned} \gamma _t = -2 \sqrt{\frac{5}{6}} B \sqrt{S_t m^*} \ge 0 \end{aligned}$$
(21)

The parameters in Eqs, 1821 are defined as follows:

$$\begin{aligned} S_n = 2 Y^* \sqrt{R^* \delta _n} \end{aligned}$$
(22)
$$\begin{aligned} S_t = 8 G^* \sqrt{R^* \delta _n} \end{aligned}$$
(23)
$$\begin{aligned} B = \frac{\ln (e)}{\sqrt{\ln ^2(e)+\pi ^2}} \end{aligned}$$
(24)
$$\begin{aligned} \frac{1}{Y^*} = \frac{1-\vartheta ^2_1}{Y_1} + \frac{1-\vartheta ^2_2}{Y_2} \end{aligned}$$
(25)
$$\begin{aligned} \frac{1}{G^*} = \frac{2(2+\vartheta _1)}{Y_1} + \frac{2(2+\vartheta _2)}{Y_2} \end{aligned}$$
(26)
$$\begin{aligned} \frac{1}{R^*}=\frac{1}{R_1}+\frac{1}{R_2} \end{aligned}$$
(27)
$$\begin{aligned} \frac{1}{m^*}=\frac{1}{m_1}+\frac{1}{m_2} \end{aligned}$$
(28)

where e is the coefficient of restitution, \(\vartheta \) is the Poisson’s ratio, Y the Young’s modulus, R the radius of a particle, m the mass of a particle and the subscripts 1 and 2 are the identifiers for two particles in contact.

Appendix 2

In this appendix we present a validation of the coupling between the fluid and solid phases. The numerical method for the fluid solver has already been validated before [see for example 43]. Also, the dispersed phase solver has been broadly used and validated in several previous investigations [16, 26, 45].

In order to carry out the validation, we perform additional simulations of a flow over a flat bed channel, different than those presented in this article. The setup reproduces the simulations carried out by Schmeeckle [45]. The configuration of the system is a rectangular channel of 0.12 m long, 0.06 m wide and 0.04 m deep. The mean velocities of the considered cases for comparisons are U = 0.2, 0.3, 0.4 and 0.5 m/s. Let us note that given these high velocities, here our algorithm works as a LES model with numerical dissipation instead of a DNS one (for this approach, see for example [33]). Following the reference case, our flow is driven by a pressure gradient, with periodic lateral boundary conditions, a free-slip rigid lid at the top, a solid wall at the bottom, and a two-way coupling approach.

Schmeeckle [45] used 115, 728 spherical particles with a diameter \(d = 0.5\) mm, which corresponds to 7.57 cm\(^3\) of solid material. We consider 28, 800 hemispherical particles stuck to the bed in order to have static particles at the bottom. In this way, we fulfill the condition used by [45] of particles with almost zero velocity at the bottom. Furthermore, we use 101, 328 spherical mobile particles. The diameter that we use is the same that the reference case. According to these conditions, the solid material of our simulations is 7.57 cm\(^3\), coinciding with the value employed by [45]. More parameters of these simulations are shown in Table 4.

Table 4 Particles parameters of comparison cases

A comparison of the bulk velocity profiles of the flow in the downstream direction is shown in Fig. 12a. The concentration of particles in the vertical direction is also compared in Fig. 12b. These results show that our model captures the region of slower flow velocity in and near the sediment bed, the development of the logarithmic velocity profile, and the distribution of particles in the vertical direction, which is directly related to the fluid-particle interactions.

Fig. 12
figure 12

Comparison of bulk velocity profiles of the flow (U) and volumetric concentration of particles (C). The horizontal axis corresponds to the non-dimensional vertical coordinate by the channel depth. Data extracted from [45] are plotted as black square, soft gray diamond, white circle and gray triangle and they represent U = 0.2, 0.3, 0.4 and 0.5 m/s respectively. Continuous lines are our simulated results. a In the original plot from [45], the bulk velocity was divided by the friction velocity. We extract it from the plot in order to directly compare the flow field. b Concentration is in volume of particles divided by volume of fluid

Additionally, we compare the velocity profile for a bulk velocity U = 0.2 m/s using a one-way and two-way coupling approach (shown in Fig. 13). For the one-way coupling case, the profile matches theoretical curves (see 29). From the perspective of the flow, this is similar to the boundary layer flow without sediments. Even though 29 can be used to calculate separately every region of a velocity profile without sediment particles [49, 54], the equation for the buffer layer can represent the whole inner layer accurately [48]. Here, the inner layer is composed by the viscous, buffer, and logarithmic layers. This equation is also known as the Spalding’s law of the wall [54] and we use it for the range \(0 \le z^+ \le 350\). We use the outer-layer equation from 29 outside this range [13].

Fig. 13
figure 13

(color online) Bulk velocity profile for U = 0.2 m/s. Red line corresponds to simulations with a one-way coupling approach, whereas the blue line corresponds to simulations with a two-way coupling approach. In black line the log-law is plotted and in dashed line the Spalding’s law is plotted [48, 49, 54]. The consideration of the two-way coupling approach yields to a completely different development of the bulk velocity profile

$$\begin{array}{llll} \frac{U}{u^*} = \left\{ \begin{array}{lll} z^+ & \text{if } \,z^+ \le 5 & {\rm Viscous sublayer} \\ z^+ - e^{-\kappa B} [e^{\kappa u^+} - 1 - \kappa u^+ -\frac{(\kappa u^+)^2}{2} - \frac{(\kappa u^+)^3}{6}] &{} \text{if }\, 5 z^+ \le 30 & {\rm Buffer layer} \\ \frac{1}{\kappa } \ln z^+ + B & \text{if }\, 30< z^+ \le 350 & \qquad {\rm Log. layer} \\ \frac{1}{\kappa } \ln z^+ + B + \frac{2 \Pi }{\kappa } \sin ^2(\frac{\pi }{2} \frac{z}{\delta }) & \text{if } \,350 < z^+ \le 1750 & {\rm Outer layer} \\ \end{array} \right. \end{array}$$
(29)

In 29, \(u^*\) is the friction velocity, \(z^+ = \frac{z u^*}{\nu }\), \(\nu \) is the kinematic viscosity, \(u^+ = \frac{U}{u^*}\), \(\delta \) is the boundary layer thickness, \(\kappa = 0.41\) is the von Kármán constant, \(\Pi = \frac{\kappa A}{2}\) and A and B are experimental coefficients with \(A = 2.5\) for flat plates and \(B = 5.0\) [14]. In the outer-layer we use the approximation \(\delta \approx 1750 \frac{\nu }{u^*}\) [54].

Rights and permissions

Reprints and permissions

About this article

Check for updates. Verify currency and authenticity via CrossMark

Cite this article

González, C., Richter, D.H., Bolster, D. et al. Characterization of bedload intermittency near the threshold of motion using a Lagrangian sediment transport model. Environ Fluid Mech 17, 111–137 (2017). https://doi.org/10.1007/s10652-016-9476-x

Download citation

  • Received:

  • Accepted:

  • Published:

  • Issue Date:

  • DOI: https://doi.org/10.1007/s10652-016-9476-x

Keywords

Navigation