You are on page 1of 23

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/257908440

3D Lagrangian modelling of saltating particles diffusion


in turbulent water flow

Article  in  Acta Geophysica · February 2014


DOI: 10.2478/s11600-012-0003-2

CITATIONS READS
35 206

3 authors, including:

Robert J. Bialik Vladimir Nikora


Polish Academy of Sciences University of Aberdeen
38 PUBLICATIONS   204 CITATIONS    190 PUBLICATIONS   6,026 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Ecological role of aquatic bryophytes View project

All content following this page was uploaded by Robert J. Bialik on 07 June 2015.

The user has requested enhancement of the downloaded file.


Acta Geophysica
vol. 60, no. 6, Dec. 2012, pp. 1639-1660
DOI: 10.2478/s11600-012-0003-2

3D Lagrangian Modelling of
Saltating Particles Diffusion in Turbulent Water Flow
Robert J. BIALIK1, Vladimir I. NIKORA2, Paweł M. ROWIŃSKI1
1
Institute of Geophysics, Polish Academy of Sciences, Warsaw, Poland
e-mails: rbialik@igf.edu.pl (corresponding author), pawelr@igf.edu.pl
2
School of Engineering, University of Aberdeen, Aberdeen, United Kingdom
e-mail: v.nikora@abdn.ac.uk

Abstract
A 3D Lagrangian model of the saltation of solid spherical particles
on the bed of an open channel flow, accounting for turbulence-induced
mechanisms, is proposed and employed as the key tool of the study. The
differences between conventional 2D models and a proposed 3D saltation
model are discussed and the advantages of the 3D model are highlighted.
Particularly, the 3D model includes a special procedure allowing genera-
tion of 3D flow velocity fields. This procedure is based on the assump-
tion that the spectra of streamwise, vertical and transverse velocity
components are known at any distance from the bed. The 3D model was
used to identify and quantify effects of turbulence on particle entrain-
ment and saltation. The analysis of particle trajectories focused on their
diffusive nature, clarifying: (i) the effect of particle mobility parameter;
(ii) the effect of bed topography; and (iii) the effect of turbulence. Spe-
cifically, the results of numerical simulations describing the above-
mentioned effects on the change in time of the variance are presented. In
addition, the change in time of the skewness and kurtosis, which are
likely to reflect the turbulence influence on the spread of particles, are
also shown. Two different diffusion regimes (local and intermediate) for
each of the investigated flow conditions are confidently identified.

Key words: Lagrangian approach, particle diffusion, particle entrain-


ment, saltating particle trajectories, sediment transport, turbulence-
particle interaction.

© 2012 Institute of Geophysics, Polish Academy of Sciences

Unauthenticated
Download Date | 6/7/15 12:57 PM
1640 R.J. BIALIK et al.

1. INTRODUCTION
Depending on the near-bed flow velocity, sediment particles on the bed may
slide, roll, or if the flow velocity exceeds some critical value, jump down-
stream. This latter form of sediment transport is called saltation. These
modes of bed-load transport were defined by Einstein (1937, 1950) who de-
veloped a stochastic theory for their description. His theory is directly
associated with the particle diffusion concept (Nikora et al. 2001), which
from a mathematical point of view defines a growth rate of the second cen-
tral moment of particle coordinates with time (note that higher order
moments are also useful to consider). During the last two-three decades, par-
ticle diffusion and the processes of sediment movement at different spatial
and time scales have received increasing attention (e.g., Drake et al. 1988,
Habersack 2001, Nikora et al. 2001, 2002, Marion et al. 2003,Wong et al.
2007, Bottacin-Busolin et al. 2008, Radice and Ballio 2008, Lukerchenko et
al. 2009b, Radice et al. 2009, 2010, Ancey 2010, Bradley et al. 2010, Ganti
et al. 2010, Tucker and Bradley 2010, and Bialik 2011b).
Nikora et al. (2001, 2002) defined three distinctly different scale ranges
of bed particle diffusion in bed-load transport: local, intermediate and global.
The local-range diffusion corresponds to short time intervals, where
X ′2 (t ) ~ t 2γ X with γx = 1.0 and Y ′2 (t ) ~ t 2γ with γy = 1.0; X ′2 and Y ′2 denote
Y

variances of particle positions, and γx, γy are the scaling exponents of particle
diffusion in streamwise X and transverse directions Y, respectively. The local
range corresponds to distances that a particle travels during a single jump.
The intermediate range of scales relates to longer time scales, involving
many jumps. Nikora et al. (2002) suggested that for the intermediate range
the particle motion may be either: (i) sub-diffusive when γx < 0.5 and
γy < 0.5; (ii) normally-diffusive when γx = 0.5 and γy = 0.5; or (iii) super-
diffusive when γx > 0.5 and γy > 0.5, depending on the dominating diffusion
mechanisms. From a physical point of view, the intermediate range corre-
sponds to durations between two successive periods of particle repose that
consist of many particle jumps. The global range of scales of saltating parti-
cle diffusion consists of many intermediate ranges, with γx < 0.5 and γy < 0.5.
Drake et al. (1988), based on the experimental data carried out in the Duck
Creek, suggested that at very long periods of time (“global range” of Nikora
et al. 2001) particle diffusion is Fickian, i.e., γx = 0.5 and γy = 0.5. However,
the reanalysis of their data set by Nikora et al. (2002) led to a different con-
clusion, supporting the sub-diffusion conjecture proposed in Nikora et al.
(2001). The factual support for the described three-range concept remains
limited, requiring more extensive studies.
The overall focus of the present paper, thus, is on the analysis of the
influence of turbulence on the entrainment and trajectories of saltating

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1641

particles in water flow from the Lagrangian perspective. Specifically, two


objectives will be pursued: (i) to provide some additional information on
scaling behavior of particle statistical moments and clarify the boundaries
between different scaling regimes proposed in Nikora et al. (2001); and
(ii) to identify and quantify the effects of turbulence, relative particle size,
and particle mobility factor on parameters of particle diffusion.
These objectives are addressed by use of the 3D model of spherical
saltating particles in which turbulence is employed explicitly. It is important
to note that the numerical models of saltation have been mostly two-
dimensional (Murphy and Hooshiari 1982, Wiberg and Smith 1985, Lee and
Hsu 1994, Niño and Garcia 1994, 1998a, Lee et al. 2000, 2002, Ancey et al.
2002, Lukerchenko et al. 2006 and Bialik 2011a). Sekine and Kikkawa
(1992) developed the particle-bed collision model by introducing a 3D sto-
chastic particle-bed collision process. Lee et al. (2006) presented the calcu-
lated three-dimensional trajectories of saltating particles in open channel
flows and developed the deterministic particle motion model. Recently,
Wang et al. (2009) has developed the 3D saltating model that is able to
simulate the continuous saltating trajectories of several particles. Lukerchen-
ko et al. (2009a) took into account the translational and rotational particle
motion and proposed a stochastic method to compute the particle-bed colli-
sion process. In spite of these improvements in the saltation modeling, much
remains to be done and this paper is a stepping stone for the further
advancement.

2. MODEL FOR SALTATING PARTICLES


2.1 The 3D model of the spherical particle motion
Our study employs a Lagrangian model involving the combination of all
fundamental forces acting on a single spherical saltating particle in a fluid
flow. The drag force, lift force, added mass force, and the submerged weight
are explicitly taken into account in this 3D model which is an extension of
the earlier proposed 2D model (Bialik 2011a). Its underlying equations are
similar to those given by Lee et al. (2006):

⎛ ⎞
du p ρ f Ad CL ( urT2 − urB2 ) ⎜ w f − wp u f − up ⎟
m = ⎜ ⎟
dt 2 ur
(u − u p ) + (v f − vp )
2 2
⎜ ⎟
⎝ f ⎠
ρ f Ad CD ur2 ⎛ u f − u p ⎞
+
2

ur
⎟ + (ρ p − ρ f ) gd 3 sin α , (1)
⎝ ⎠

Unauthenticated
Download Date | 6/7/15 12:57 PM
1642 R.J. BIALIK et al.

⎛ ⎞
dv p ρ f Ad CL ( urT2 − urB2 ) ⎜ w f − wp v f − vp ⎟
m = ⎜ ⎟
dt 2 ur
(u − u p ) + (v f − vp )
2 2
⎜ ⎟
⎝ f ⎠
ρ f Ad CD ur2 ⎛ v f − v p ⎞
+ ⎜ ⎟, (2)
2 ⎝ ur ⎠

⎛ ⎞
dwp ρ f Ad CL ( urT2 − urB2 ) ⎜ w f − wp w f − wp ⎟
m = ⎜ ⎟
dt 2 ur
(u − u p ) + (v f − vp )
2 2
⎜ ⎟
⎝ f ⎠
ρ f Ad CD ur2 ⎛ w f − wp ⎞

2

ur
⎟ − (ρ p − ρ f ) gd 3 cos α , (3)
⎝ ⎠
dx p dy p dz p
= up , = vp , = wp , (4)
dt dt dt
where up , vp , wp denote streamwise, transverse and vertical components of
particle velocity, respectively; and uf , vf , wf represent streamwise, trans-
verse and vertical components of fluid velocity, respectively; the coordinates
xp , yp , zp describe particle position; CD denotes the drag coefficient and CL
stands for the lift coefficient; m = (ρp + Cmρf)πd 3/6, where Cm = 0.5 is the vir-
tual mass (added mass) coefficient, ρp and ρf are the particle and fluid densi-
ties; ur = [(up – uf)2 + (wp – wf)2 + (vp – vf)2 ]1/2 is the relative velocity of
particle and fluid; subscripts T and B refer to the top and the bottom of the
moving particle, respectively; Ad is the particle cross-sectional area; d is the
particle diameter; t is the saltation time; α is the angle of channel bed; and
g is gravity acceleration. The drag coefficient is calculated with use of the
empirical formula proposed by Yen (1992):

CD =
24
Re p
(
1 + 0.15 Re p + 0.017 Re p ) −
0.208
1 + 104 Re −p0.5
, (5)

where Rep = |ur| d/ν denotes the particle Reynolds number; and ν is kine-
matic viscosity. The lift coefficient is obtained with use of the formula given
by Lee at al. (2006):
CL = 1.45 × 10 −5 × Re1.7325
p . (6)

2.2 Collision with the channel bed


To obtain the continuous, saltating particle trajectory, the bed-collision
process has to be included in the model. In our study we used the hard

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1643

sphere model, based on the impulsive equations proposed by Crowe et al.


(1998):
u p|out = u p|in − ( sin θ cos ϕ + f sin ϕ ) ( u p|in sin θ cos ϕ
1 + e
+ v p|in sin θ sin ϕ + wp|in cos θ ) , (7)
1 + (d D)
3

v p|out = v p|in − ( sin θ sin ϕ − f cos ϕ ) ( u p|in sin θ cos ϕ


1 + e
+ v p|in sin θ sin ϕ + wp|in cos θ ) , (8)
1 + (d D)
3

wp|out = wp|in − cos θ ( u p|in sin θ cos ϕ


1+ e
+ v p|in sin θ sin ϕ + wp|in cos θ ) , (9)
1 + (d D)
3

where up , vp , wp denote streamwise, transverse and vertical components of


particle velocity, respectively. The subscript “in” means values before colli-
sion, while subscript “out” denotes values after collision; d is the diameter of
the moving particle and D is the diameter of the particle which lays on the
bed; parameters e and f represent the restitution and friction coefficients,
respectively and are taken as in Niño and Garcia (1998b), i.e., e = 0.75 –
0.25τ∗/τ∗c and π/6 < θ < π/3 measured in the direction of the movement of
saltating particle and 0 < ϕ < 2π. These two angles are random variables rep-
resenting the irregularity of collision or packing of other grains on the chan-
nel bottom and their values are obtained from the uniform probability
density function and with additional assumption that the shadow effect may
be taken into particle-bed collision process account. This effect is included
in simulations by analogy to the 2D model of collisions of non-uniform
saltating grains proposed by Bialik (2011b).

2.3 Model for flow velocity field


We use a turbulence generator which is relatively simple and fast compared
to the commonly used simulation procedures such as LES or DNS. A gen-
erator, initially proposed and employed by Nikora et al. (2001) and later
extensively used by Bialik et al. (2010) and Bialik (2011a), has been modi-
fied to generate 3D flow velocity field. In this method it is assumed that the
wave-number auto-spectra of velocity components are known at any distance
from the bed. The main differences with the previous applications are that
the vertical and longitudinal components of velocity fluctuations are now
correlated, simulating the experimentally observed correlations, based on the
degree of similarity of turbulence, and that the transverse component of

Unauthenticated
Download Date | 6/7/15 12:57 PM
1644 R.J. BIALIK et al.

velocity fluctuations is generated based on the continuity equation. Our pro-


cedure is briefly described below.
Assuming that the instantaneous streamwise, transverse and vertical
components of flow velocity vector are decomposed into mean flow velocity
u f , v f , w f and velocity fluctuations u ′ , v′ and w′ , that v f , w f are negligi-
ble and the logarithmic law holds in the near-bed flow region, we get
u f = u f + u′ , v f = v′ , and w f = w′ . (10)

The origin of the vertical coordinate for the velocity distribution is


assumed to be at the top of bed particles with the mean streamwise velocity
below this level assumed to be equal to zero (Fig. 1).
The time series of velocity fluctuations in streamwise and vertical direc-
tions are deduced from the assumed wave-number auto-spectra of velocity
components and the vertical distributions of the relative turbulence intensi-
ties which for open channel flow may be described by Nezu and Nakagawa’s
(1993) expressions. As for wave-number auto-spectra, it is assumed that they
consist of four specific ranges (Nikora 1999, 2005, and Nikora and Goring
2000): (i) the production range; (ii) the intermediate range where energy
production and energy transfer to smaller scales co-exist; (iii) the inertial
subrange; and (iv) the dissipative range. Knowing the shapes of the auto-
spectra, realizations of the surrogate velocity fields can be obtained using
Inverse Fast Fourier Transform.
As shown in Fig. 1, the flow is divided into layers of dz = d/2 thickness
and velocity time series are generated for each layer separately, by employ-
ing the Monte Carlo approach. The transverse correlation between velocity
time series is simulated using the same phases of spectral components of the
velocity signals across the flow. The phases for particular spectral compo-
nents are defined using a random number generator for a uniformly distrib-
uted variable (see Bialik et al. 2010 for details). The vertical and
longitudinal components of velocity fluctuations are generated with the

Fig. 1. Levels at which velocity time series are generated (adapted from Bialik et al.
2010).

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1645

Fig. 2. An example of quasi-Lagrangian fluctuation velocity time series of flow


velocity components.

additional assumption that the correlation between them in the near-bed


region satisfies the experimental data (e.g., Nezu and Nakagawa 1993):
uw
0.4 < R = − < 0.5 . (11)
u ′w′
After the generation of the streamwise velocity time series, the vertical
velocity component has been generated by the trials and errors approach
until Eq. (11) is satisfied.
Once velocity fluctuations u ′ and w′ are obtained, we can find v′ from
the continuity equation
du f dv f dwf du′ dv′ dw′
= = = 0 and + + = 0 (12)
dx dy dz dx dy dz
as
yi +1
⎛ du ′ dw′ ⎞
v′( x, yi +1 , z, t ) = v′( x, yi , z, t ) − ∫ ⎜⎝ dx ( x, y, z, t )
yi
+
dz
( x, y, z, t ) ⎟ dy .

(13)

To obtain fluid velocities involved in Eqs. (1)-(4) we sampled flow


velocities from simulated velocity fields at instantaneous particle positions.
We define these flow velocities as quasi-Lagrangian velocities, to highlight
their relevance to the instantaneous Lagrangian particle velocities. Figure 2
presents an example of quasi-Lagrangian turbulent time series of flow veloc-
ity components obtained with use of the presented method.

3. RESULTS
3.1 Numerical simulation details
The system of equations (1)-(4) is numerically solved using the fourth-order
Runge–Kutta method. The initial conditions are taken as up(t0) = 0,

Unauthenticated
Download Date | 6/7/15 12:57 PM
1646 R.J. BIALIK et al.

vp(t0) = 0, wp(t0) = 0, x(t0) = 0, and z(t0) = 0.5d. These conditions mean that
the saltating particle is on the bed and does not move (Fig. 3). Such initial
conditions differ from the usually used Abbott and Francis’ (1977)
conditions up(t0) = 2u∗, wp(t0) = 2u∗ and allow us to analyze the incipient
motion of a particle. The entrainment in our model occurs as a result of
velocity fluctuations which directly influence the drag on a particle and the
lift force (Fig. 4).
To explore and verify the hypotheses presented in the introduction to
this paper, the simulations were completed for the following conditions:
(i) the relative sizes of moving particles are selected to be d/D = 0.5, 1, and 2
(Fig. 5); and (ii) for each relative size of moving particles, the simulations
were carried out for three values of particle mobility parameter K = 1.1, 1.5,
and 2.5. The simulation scenarios are briefly summarized in Table 1.

Fig. 3. Particle entrainment in the situation when the particle is on top of bed.

Fig. 4. Scheme of the initiation of the particle movement by velocity fluctuations.

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1647

Fig. 5. The relative sizes of moving particles (a) d/D = 2; (b) d/D = 1; (c) d/D = 0.5.

Table 1
Simulation matrix
Test series 1 2 3 4 5 6 7 8 9
K 1.1 1.5 2.5 1.1 1.5 2.5 1.1 1.5 2.5
d [mm] 2 2 2 2 2 2 2 2 2
d/D 2 2 2 1 1 1 0.5 0.5 0.5
σ' 0; 0.3σ ; 0.7σ ; σ

In Table 1, D denotes the size of the particles on the bed, d stands for the
diameter of the saltating particle, σ′ is the turbulence intensity used in simu-
lations, σ is the turbulence intensity from Nezu and Nakagawa’s (1993) rela-
tionships, and K is the mobility parameter which is defined as
1 ρ u*2
K = , (14)
θ c ( ρ s − ρ ) gd

where θc = 0.05 is the critical value of the bed shear stress. The shear veloci-
ty u∗ required for generation of the turbulent velocity time series is obtained
based on the value of the mobility parameter.

3.2 Validation of the model


The experimental data of Niño and Garcia (1998b) were used to validate the
saltating particle model. They are compared with the results of simulations
calculated according to both 3D and earlier developed 2D models. The 2D
saltation model proposed by Bialik (2011a) is used to obtain the 2D particle
trajectory, mean dimensionless saltation length and height for comparison
purposes.
Figure 6 shows a comparison of calculated mean dimensionless saltation
height, Hs, and length, Ls, with the experimental results of Niño and Garcia
(1998b) for a particle with diameter d = 0.53 mm (vertical bars show
plus/minus two standard deviations; the horizontal axis shows the normal-
ized bed shear stress). It is important to stress that the dimensionless mean
values of the saltation length and height, obtained using the 3D model, agree
very well with the experimental data. The 2D model leads to the overestima-
tion of the saltation length and height. Lukerchenko et al. (2009a) obtained

Unauthenticated
Download Date | 6/7/15 12:57 PM
1648 R.J. BIALIK et al.

Fig. 6. Comparison of numerical simulations with use of 2D and 3D model with


experimental data of Niño and Garcia (1998b) for particles with d = 0.53 mm.

similar results using the 3D saltation model with rotation. They explained
their result by the transformation of the impulse force during the particle-bed
collisions in all three directions, which is impossible to achieve using the 2D
model.

3.3 Particles’ entrainment and deposition


Let us employ the definition of the particle entrainment suggested by Drake
et al. (1988) who proposed that: “Because some particles in repose vibrated
or jostled against their neighbors without going anywhere, we defined
entrainment operationally as continuous movement a net horizontal
distance of one particle diameter.” Figures 7 and 8 show examples of
particles’ positions in the x-t plane during the particle movement initiation.
We can notice that particles start to move at different moments of time,
reflecting, most likely, specific hydrodynamic conditions at the moment of
entrainment.
Note that the values of velocity components of a moving particle before
the collision are a result of the system of ordinary differential equations
(Eqs. (1)-(4)) and they strongly influence the velocities after collision
through Eqs. (7)-(9). The velocities after collision may eventually become
low enough that those particles are assumed to stay on the channel bed, in
other words, to be in a depositional mode. Otherwise they move either in a
saltation mode or are entrained if the assumed boundary of entrainment is
crossed (see Fig. 7). As shown in Fig. 9, the particle may be stopped for
longer period of time after some displacement and then it may be entrained
again if drag and lift force exceed threshold values. The latter result mostly
highlights the fact that the presented model should be able to reproduce the
global range of particle diffusion as the detrainment effect is taken into
account.

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1649

Fig. 7. Examples of particles’ positions X(t)/d at movement initiation.

Fig. 8. Time between initiation of particle movement (i.e., vibration) and entrain-
ment for one specific trajectory of a particle with d = 2 mm.

Fig. 9. Example of particles’ deposition and rest period.

Unauthenticated
Download Date | 6/7/15 12:57 PM
1650 R.J. BIALIK et al.

3.4 Diffusion of saltating particles


Figures 10 and 11 show the examples of temporal changes of particles’ loca-
tions for two cases, with and without turbulence included. The diffusion
phenomenon can be easily noticed in both figures. However, if turbulence is
included in the simulations, the diffusion effect becomes much stronger for
the X coordinate, remaining almost the same for the Y coordinate. It is likely
that particle collisions with the bed play the dominant role in the transverse
diffusion, well exceeding the effect of turbulence.
As was briefly described in the Introduction, the variance of particle co-
ordinates ( X ′2 , Y ′2 ) increase in time as ( X ′2 (t ) ~ t 2γ X , Y ′2 (t ) ~ t 2γ Y ). For com-
paring the results obtained for different transport conditions, the change in
time of the second-order moments of particle coordinates will be presented
in the dimensionless form as a function of the normalized time tu∗/d, where d
is the diameter of the moving particle, t is the time of saltation, and

Fig. 10. Particle trajectories (a) X(t)/d and (b) Y(t)/d for K = 2.5 and d/D = 1 without
inclusion of turbulence (σ′ = 0).

Fig. 11. Particle trajectories (a) X(t)/d and (b) Y(t)/d for K = 2.5 and d/D = 1 with
inclusion of turbulence (σ′ = σ).

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1651

u∗ is the shear velocity. Thus, the following relationships will be employed


(Nikora et al. 2002):
2γ X 2γ Y
X ′2 ⎛ tu ⎞ Y ′2 ⎛ tu ⎞
2
= ⎜ *⎟ and 2
= ⎜ *⎟ , (15)
d ⎝ d ⎠ d ⎝ d ⎠
where γX and γY stand for the diffusion scaling exponents that for the local
range should be γX = γY = 1.0, and for the global range they are expected to
be γX , γY < 0.5 (Nikora et al. 2001, 2002). In the following subsections we
will summarize the results related to the effects of particle mobility parame-
ter, turbulence, and relative particle size. The turbulence is quantified
in terms of turbulence intensity σ′ using the value σ from Nezu and Nak-
agawa’s (1993) relationships as a measure (i.e., the range σ′ = 0…σ has
been explored).

3.4.1 Near-field effect


In general, there may be several ways of how to choose the starting points of
particle trajectories for the diffusion analysis. For example, Nikora et al.
(2001, 2002) used the randomly chosen collision points when saltation was

Fig. 12. Change in time of the second-order moments of particle positions for
d/D = 2: (a) X′2(t) and (b) Y′2(t); E denotes entrainment as starting points of particle
trajectories.

Unauthenticated
Download Date | 6/7/15 12:57 PM
1652 R.J. BIALIK et al.

already well developed, i.e., the initial entrainment conditions were “forgot-
ten” by moving particles. Alternatively, the starting points of particle trajec-
tories can be chosen as entrainment points. In this latter case, the particle
behavior at small diffusion times may be strongly biased by the particular
entrainment mechanisms or entrainment conditions, i.e., the scaling diffusion
relationships at small diffusion times may be different from those in Nikora
et al.’s (2001, 2002) approach. To illustrate potential differences between
these approaches, Fig. 12 shows the scaling plots for both cases. As one can
see, for (tu∗)/d > 3 these approaches result in identical scaling behavior of
particle trajectories revealing local, intermediate and global ranges. Howev-
er, at smaller (tu∗)/d the plots differ from each other revealing much higher
scaling exponents when the starting trajectory points are defined at entrain-
ment points, compared to expected “ballistic” value of approximately 1 as in
Nikora et al. (2001, 2002). Clearly, this difference is due to the entrainment
effect which represents, essentially, some-kind of a “near-field” effect.
In our considerations below we use the second approach as it may provide
some additional information on entrainment effects on particle trajectories
shortly after particle detachment from the bed, typically observable at
(tu∗)/d < 3.

Fig. 13. Change in time of the second-order moments of particle positions for
d/D = 2: (a) X′ 2(t) and (b) Y′ 2(t).

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1653

Fig. 14. Change in time of the second-order moments of particle positions for
d/D = 1: (a) X′ 2(t) and (b) Y′ 2(t).

3.4.2 Effect of particle mobility parameter


Figures 13 and 14 show the change in time of the variance of particle posi-
tions for two different relative sizes of particles for a range of values of the
particle mobility parameter K. One may observe that the data for different K
collapse around single curves highlighting the independence of the diffusion
scaling on this parameter once the data are presented in the form of Eq. (15).
Also, for both relative particle sizes three different diffusion regimes can
be identified for the X coordinate: (i) local, for (tu∗)/d < 15 with γX ~ 1;
(ii) intermediate, for 15 < (tu∗)/d < 800 with γX ~ 0.9; and (iii) global for
800 < (tu∗)/d with γX ~ 0.33. These results suggest that the spread of
saltating particles in the streamwise direction is super-diffusive within the
local and intermediate ranges and subdifussive within the global range.
However, for the Y coordinate there are only two diffusion regimes
observed: (i) local, for (tu∗)/d < 15 with γX ~ 1.1, that is also super-diffusive
as for the X coordinate; and (ii) intermediate, for (tu∗)/d > 15 with γX ~ 0.55,
that is close to normal. The absence of the clearly defined global range for
the Y coordinate can be explained by insufficient statistics (very long simula-
tions would be required to resolve this range unambiguously). It is also in-
teresting to note that the diffusion of saltating particles for local range is

Unauthenticated
Download Date | 6/7/15 12:57 PM
1654 R.J. BIALIK et al.

isotropic while for intermediate and global is anisotropic (γX ≠ γY) and that
the boundary between the local and intermediate ranges of scales is the same
for X and Y coordinates

3.4.3 Effect of turbulence


The effect of turbulence can be illustrated using simulations for the particle
mobility parameter K = 2.5 and the relative particle size d/D = 1, as shown
in Fig. 15. These results are based on the simulations of saltation for a range
of turbulence intensity from zero to σ. One may note that in the streamwise
direction the spread of particles noticeably increases with the turbulence
intensity. We can also identify three diffusion regimes with a clear tendency
to slow down the diffusion process with its duration. However, for the
Y coordinate no turbulence effect can be observed, i.e., all data collapse on
single curve where only two diffusion ranges can be observed: (i) local with
γY ~ 1.03, and (ii) intermediate with γY ~ 0.53.
The results for the higher (third Sk = E(x – µ)3/σ3 and fourth order
Ku = E(x – µ)4/σ4 – 3) moments of particle positions are presented in Figs. 16
and 17. It is easy to notice that for the small (tu∗)/d < 15, the values of these

Fig. 15. Change in time of the second-order moments of particle positions for
K = 2.5 and d/D = 1 and for a range of turbulence intensities: (a) X′ 2(t) and (b) Y′ 2(t).

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1655

Fig. 16. Change in time of the normalized third-order moments of particle positions
for a range of turbulence intensities: (a) SkX and (b) SkY.

Fig. 17. Change in time of the normalized fourth-order moments of particle positions
for a range of turbulence intensities: (a) KuX and (b) KuY.

two moments deviate from the Gaussian distribution and tend to zero for
(tu∗)/d > 15. The high values of the skewness and kurtosis at small (tu∗)/d
reflect high level of intermittency of particle trajectories, in agreement with
Nikora et al. (2002). This intermittency is most likely the reason of high val-
ues of scaling exponents obtained for the near-field effect range of scales.
The simulations with turbulence switched off show zero skewness and kur-
tosis as one would expect for the Gaussian distribution for any given diffu-
sion time. This result suggests that turbulence plays a significant role in the
initiation of particles’ movement.

3.4.4 Effect of relative size of particles


Figure 18 summarizes the simulation results obtained for a range of relative
particle sizes, from d/D = 0.5 to d/D = 2, at the same particle mobility para-
meter K = 2.5 and the same value of turbulence intensity for each considered

Unauthenticated
Download Date | 6/7/15 12:57 PM
1656 R.J. BIALIK et al.

Fig. 18. Change in time of the second-order moments of particle positions for
a range of relative particle sizes and K = 2.5: (a) X′ 2(t) and (b) Y′ 2(t).

case σ′ = σ. For the X coordinate, we can identify three ranges of scales:


(i) local, with γX ~ 1.07; (ii) intermediate, with γX ~ 0.87; and (iii) global,
with γX ~ 0.32. These results are almost the same as presented in Figs. 13a
and 14a. We can also observe that the particle diffusion is only very weakly
dependent on the relative particle size. On the other hand, for the Y coordi-
nate there is a much stronger effect of the relative particle size which reflects
a strong positive correlation between diffusion and D/d. We can also identi-
fy, similar to Figs. 13b and 14b, two ranges of scales: (i) local, with
γY ~ 1.32; and (ii) intermediate, with γY ~ 0.5. This result may be explained
by the transformation of the momentum during the particle-bed collision in
the three directions and by the fact that the particle reaction to fluid velocity
in streamwise direction is much faster than in the transverse direction.

4. CONCLUDING REMARKS
This paper reports the results of numerical simulations of saltating particles’
movement and their diffusion in turbulent water flows. Firstly, we confirmed
the conceptual model for diffusion of moving particles proposed by Nikora
et al. (2001, 2002), that consists of at least three ranges of scales, i.e., local,

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1657

intermediate and global. Secondly, the results of our simulations, based on a


more advanced turbulence generator and particle motion model (that allows
entrainment and deposition), and employing wider ranges of controlling fac-
tors, provide new information extending the previous results. Specifically,
the main additions to the previous studies relate to the following: (i) the
spread of particles in the transverse direction reveal two ranges of scales
(i.e., local and intermediate) compared to three ranges for the longitudinal
direction; (ii) the spread of the particles in the longitudinal direction mostly
depends on the turbulence, whereas in transverse direction it depends more
on the relative particle size; and (iii) the boundary between the intermediate
range and the global range is much higher than (tu∗)/d ~ 15 suggested in
Nikora et al. (2002) from a comparison of Drake et al.’s (1988) field data
with simulations, being approximately 800. The latter result highlights the
fact that the boundaries between diffusion regimes depend also on additional
parameters governing particles’ movement and the next challenge should be
to find the most suitable approach describing sediment particles diffusion for
a wide range of conditions.
We believe that the present paper represents an important step forward in
the challenging field of modeling of sediment transport in turbulent water
flows and should be helpful in the advancement of the fundamental under-
standing of particle diffusion and transport in open-channel flows.

A c k n o w l e d g m e n t s . The work presented here was supported by The


Polish National Science Centre Grant No. N N306 658140 “Development of
two-layers hydrodynamic model of bed-load sediment transport comprising
saltation motion and moving bed”, and by EPSRC, UK (EP/G056404/1)
within the project “High resolution numerical and experimental studies of
turbulence-induced sediment erosion and near-bed transport.”

References

Abbott, J.E., and J.R.D. Francis (1977), Saltation and suspension trajectories of solid
grains in a water stream, Phil. Trans. R. Soc. Lond. A 284, 1321, 225-254,
DOI: 10.1098/rsta.1977.0009.
Ancey, C. (2010), Stochastic modeling in sediment dynamics: Exner equation for
planar bed incipient bed load transport conditions, J. Geophys. Res. 115,
F00A11, DOI: 10.1029/2009JF001260.
Ancey, C., F. Bigillon, P. Frey, J. Lanier, and R. Ducret (2002), Saltating motion
of a bead in a rapid water stream, Phys. Rev. E 66, 036306, DOI: 10.1103/
PhysRevE.66.036306.

Unauthenticated
Download Date | 6/7/15 12:57 PM
1658 R.J. BIALIK et al.

Bialik, R.J. (2011a), Particle-particle collision in Lagrangian modelling of saltating


grains, J. Hydraul. Res. 49, 1, 23-31, DOI: 10.1080/00221686.2010.543778.
Bialik, R.J. (2011b), Numerical study of saltation of non-uniform grains, J. Hydraul.
Res. 49, 5, 697-701, DOI: 10.1080/00221686.2011.598025.
Bialik, R.J., V.I. Nikora, P.M. Rowiński, and W. Czernuszenko (2010), A numerical
study of turbulence influence on saltating grains. In: A. Dittrich, K. Koll,
J. Aberle, and H. Geisenhainer (eds.), Proc. Int. Conf. on River Flow 2010,
08-10.09.2010, Braunschweig, Germany, Bundesanstalt für Wasserbau,
Karlsruhe, 105-112.
Bottacin-Busolin, A., S.J. Tait, A. Marion, A. Chegini, and M. Tregnaghi (2008),
Probabilistic description of grain resistance from simultaneous flow field
and grain motion measurements, Water Resour. Res. 44, W09419, DOI:
10.1029/2007WR006224.
Bradley, D.N., G.E. Tucker, and D.A. Benson (2010), Fractional dispersion in
a sand bed river, J. Geophys. Res. 115, F00A09, DOI: 10.1029/
2009JF001268.
Crowe, C., M. Sommerfeld, and Y. Tsuji (1998), Multiphase Flows with Droplets
and Particles, CRC Press, Boca Raton.
Drake, T.G., R.L. Shreve, W.E. Dietrich, P.J. Whiting, and L.B. Leopold (1988),
Bedload transport of fine gravel observed by motion-picture photography,
J. Fluid Mech. 192, 193-217, DOI: 10.1017/S0022112088001831.
Einstein, H.A. (1937), Der Geschiebebetrieb als Wahrscheinlichkeitsproblem, Ver-
lag Rascher, Zurich.
Einstein, H.A. (1950), The Bed-Load Function for Sediment Transportation in Open
Channel Flows, Tech. Bull. 1026, U.S. Dept. of Agriculture, Washington
D.C.
Ganti, V., M.M. Meerschaert, E. Foufoula-Georgiou, E. Viparelli, and G. Parker
(2010), Normal and anomalous diffusion of gravel tracer particles in rivers,
J. Geophys. Res. 115, F00A12, DOI: 10.1029/2008JF001222.
Habersack, H.M. (2001), Radio-tracking gravel particles in a large braided river in
New Zealand: a field test of the stochastic theory of bed load transport pro-
posed by Einstein, Hydrol. Process. 15, 3, 377-391, DOI: 10.1002/hyp.147.
Lee, H.-Y., and I.-S. Hsu (1994), Investigation of saltating particle motion.
J. Hydraul. Eng. 120, 7, 831-845, DOI: 10.1061/(ASCE)0733-9429(1994)
120:7(831).
Lee, H.-Y., Y.-H. Chen, J.-Y. You, and Y.-T. Lin (2000), Investigations of continu-
ous bed load saltating process, J. Hydraul. Eng. 126, 9, 691-700, DOI:
10.1061/(ASCE)0733-9429(2000)126:9(691).
Lee, H.-Y., J.-Y. You, and Y.-T. Lin (2002), Continuous saltating process of
multiple sediment particles, J. Hydraul. Eng. 128, 4, 443-450, DOI:
10.1061/(ASCE)0733-9429(2002)128:4(443).

Unauthenticated
Download Date | 6/7/15 12:57 PM
SALTATING PARTICLE DIFFUSION IN TURBULENT WATER FLOW 1659

Lee, H.-Y., Y.-T. Lin, J.-Y. You, and H.-W. Wang (2006), On three-dimensional
continuous saltating process of sediment particles near the channel bed,
J. Hydraul. Res. 44, 3, 374-389, DOI: 10.1080/00221686.2006.9521689.
Lukerchenko, N., Z. Chara, and P. Vlasak (2006), 2D numerical model of particle-
bed collision in fluid-particle flow over bed, J. Hydraul. Res. 44, 1, 70-78,
DOI: 10.1080/00221686.2006.9521662.
Lukerchenko, N., S. Piatsevich, Z. Chara, and P. Vlasak (2009a), 3D numerical
model of the spherical particle saltation in a channel with a rough fixed bed,
J. Hydrol. Hydromech. 57, 2, 100-112, DOI: 10.2478/v10098-009-0009-x.
Lukerchenko, N., S. Piatsevich, Z. Chara, and P. Vlasak (2009b), Numerical model
of spherical particle saltation in a channel with a transversely tilted rough
bed, J. Hydrol. Hydromech. 57, 3, 182-190, DOI: 10.2478/v10098-009-
0017-x.
Marion, A., S.J. Tait, and I. McEwan (2003), Analysis of small-scale gravel bed
topography during armoring, Water Resour. Res. 39, 12, 1334, DOI:
10.1029/2003WR002367.
Murphy, P.J., and H. Hooshiari (1982), Saltation in water dynamics, J. Hydraul.
Eng. 108, 11, 1251-1267.
Nezu, I., and H. Nakagawa (1993), Turbulence in Open-Channel Flows, Balkema,
Rotterdam.
Nikora, V. (1999), Origin of the “–1” spectral law in wall-bounded turbulence, Phys.
Rev. Lett. 83, 4, 734-736, DOI: 10.1103/PhysRevLett.83.734.
Nikora, V. (2005), Flow turbulence over mobile gravel-bed: spectral scaling and co-
herent structures, Acta Geophys. Pol. 53, 4, 539-552.
Nikora, V., and D. Goring (2000), Flow turbulence over fixed and weakly mobile
gravel beds, J. Hydraul. Eng. 126, 9, 679-690, DOI: 10.1061/(ASCE)0733-
9429(2000)126:9(679).
Nikora, V., J. Heald, D. Goring, and I. McEwan (2001), Diffusion of saltating parti-
cles unidirectional water flow over a rough granular bed, J. Phys. A: Math.
Gen. 34, L743, DOI: 10.1088/0305-4470/34/50/103.
Nikora, V., H. Habersack, T. Huber, and I. McEwan (2002), On bed particle diffu-
sion in gravel bed flows under weak bed load transport, Water Resour. Res.
38, 6, DOI: 10.1029/2001WR000513.
Niño, Y., and M. Garcia (1994), Gravel saltation: 2. Modeling, Water Resour. Res.
30, 6, 1915-1924, DOI: 10.1029/94WR00534.
Niño, Y., and M. Garcia (1998a), Using Lagrangian particle saltation observations
for bedload sediment transport modelling, Hydrol. Process. 12, 8,
1197-1218, DOI: 10.1002/(SICI)1099-1085(19980630)12:8<1197::AID-
HYP612>3.0.CO;2-U.
Niño, Y., and M. Garcia (1998b), Experiments on saltation of sand in water,
J. Hydraul. Eng. 124, 10, 1014-1025, DOI: 10.1061/(ASCE)0733-
9429(1998)124:10(1014).

Unauthenticated
Download Date | 6/7/15 12:57 PM
1660 R.J. BIALIK et al.

Radice, A., and F. Ballio (2008), Double-average characteristics of sediment motion


in one-dimensional bed load, Acta Geophys. 56, 3, 654-668, DOI: 10.2478/
s11600-008-0015-0.
Radice, A., F. Ballio, and V. Nikora (2009), On statistical properties of bed load
sediment concentration, Water Resour. Res. 45, W06501, DOI: 10.1029/
2008WR007192.
Radice, A., F. Ballio, and V. Nikora (2010), Statistics and characteristic scales for
bed load in a channel flow with sidewall effects, Acta Geophys. 58, 6,
1072-1093, DOI: 10.2478/s11600-010-0020-y.
Sekine, M., and H. Kikkawa (1992), Mechanics of saltating grains. II, J. Hydraulic
Eng. 118, 4, 536-558.
Tucker, G.E., and D.N. Bradley (2010), Trouble with diffusion: Reassessing hill-
slope erosion laws with a particle-based model, J. Geophys. Res. 115,
F00A10, DOI: 10.1029/2009JF001264.
Wang, H.W., H.Y. Lee, and P.N. Lee (2009), Three-dimensional saltating process of
multiple sediment particles, Int. J. Sediment Res. 24, 1, 16-32, DOI:
10.1016/S1001-6279(09)60013-5.
Wiberg, P.L., and J.D. Smith (1985), A theoretical model for saltating grains in wa-
ter, J. Geophys. Res. 90, C4, 7341-7354, DOI: 10.1029/JC090iC04p07341.
Wong, M., G. Parker, P. DeVries, T.M. Brown, and S.J. Burges (2007), Experiments
on dispersion of tracer stones under lower‐regime plane‐bed equilibrium
bed load transport, Water Resour. Res. 43, W03440, DOI: 10.1029/
2006WR005172.
Yen, B.C. (1992), Sediment fall velocity in oscillating flow, Water Resour. Environ.
Eng. Res. Rep. 11, Dept. of Civil Eng. University of Virginia.

Received 11 May 2011


Received in revised form 25 November 2011
Accepted 2 January 2012

Unauthenticated
Download Date | 6/7/15 12:57 PM
View publication stats

You might also like