You are on page 1of 5

70

sediment transport and morphodynamics


itself evolved many decades afterward and is now used for
river, coastal, and estuarine problems (e.g., Parker and Garca
2006). This notwithstanding, Exner deserves recognition as
the founder of morphodynamics (Parker 2005).
The eld of morphodynamics consists of the class of
problems for which the ow over a bed interacts strongly
with the shape of the bed, both of which evolve in time.
An introduction to morphodynamics of rivers and turbidity
currents is given in Section 2.11 of this chapter. The Exner
equation is generalized to the case of sediment mixtures in
Chapter 3.
2.6.4 Bed Load Transport Relations

Fig. 2-32.

Denition diagram for sediment mass conservation.

A large number of bed load relations can be expressed in the


general dimensionless form

the datum is given by H. The mass balance equation translates to

Rs 1 L p

uut  H ds dn  R q
R  q

bs s  qbs sds

dn
ds (2-88a)

bn n  qbn ndn

q*  q* T* , Rep , R

q* 

Or upon reduction with Eqs. (2-84) and (2-87), Eq. (2-88a)


takes the nal form
q
q

 bs  bn  vs cb  Es
t
s
n

(2-88b)

Bed level changes with time t due to bed load transport,


sediment entrainment into suspension, and sediment deposition onto the bed can be predicted with this partial differential equation. To solve this equation, it is necessary to have
relations to compute bed load transport (i.e., qbs and qbn),
_
near-bed suspended sediment concentration cband sediment
entrainment into suspension Es (Garca and Parker 1991;
Garca 2001). The basic form of Eq. (2-88b), without the
suspended sediment component, was rst proposed for the
case of a one-dimensional ow interacting with a sedimentcovered bed by Felix Exner (1925).
Felix Maria Exner was an Austrian researcher who was
active in the early part of the 20th Century. His main area
of interest was meteorology. At some point he became interested in the formation of dunes in rivers (see Leliavsky 1966).
In the course of his research on the subject, he derived and
employed one version of the various statements of conservation of bed sediment that are now referred to as Exner equations. In addition, he made an important early contribution
to one-dimensional nonlinear wave dynamics (Exner 1920).
Felix Exner was the rst researcher to state a morphodynamic
problem in quantitative terms. The term morphodynamics

(2-89)

Here, q* is a dimensionless bed load transport rate known


as the Einstein bed load number, rst introduced by Hans
Albert Einstein in 1950, and given by

 Rs  Dr  Er ds dn

1  p 

qb
D gRD

(2-90a)

where qb is the volumetric bedload transport rate, g is the


acceleration of gravity, R  (Rs  R)  R is the submerged
specic gravity of the sediment, D is the particle diameter,
and Rep  gRDD N is the particle Reynolds number.
Einsteins bed load transport model can be expressed in
dimensionless form as follows,
q*  E*p L*s

(2-90b)

with
Eb* 

Eb
gRD

(2-90c)

and
L*s 

Ls
D

(2-90d)

In these relations Eb  volumetric rate of sediment


entrainment per unit area; and Ls  particle step length (i.e.,
the particle travel distance from entrainment to deposition).
In a study of sand bed instability, Nakagawa and
Tsujimoto (1980) found that the dimensionless entrainment
rate E*b is a function of the Shields parameter T* and tried to
develop a probabilistic model for the dimensionless particle

bed load transport


L*s

step length but did not nd a simple way to characterize


this parameter (Tsujimoto and Nakagawa 1983). Sechet and
Le Guennec (1999) found experimentally that the particle
step length is related to the bursting phenomenon (i.e., ejections and sweeps) which is in agreement with observations
of near-wall particle-turbulence interactions made earlier
by Sumer and Deigaard (1981), Garca et al. (1996), and
Nio and Garca (1996). This might explain some of the difculties encountered in trying to characterize Einsteins step
length. Recently, Sumer et al (2003) were able to observe
directly the inuence of turbulence on bed load transport of
sand (D50  0.22 mm) with a set of carefully conducted lab
experiments. The Shields parameter together with the rootmean-square value (RMS) of the streamwise velocity uctuations, were correlated with the sediment transport rate.
They found that the sediment transport rate increases markedly with increasing turbulence levels. A few years earlier,
Drake et al (1988) observed a similar inuence of the ow
turbulence while observing bed load transport of ne gravel
with motion-picture photography.
Recently, Wong and Parker (2006a) conducted a tracer
study involving transport of uniform-size gravel in a large
ume at St. Anthony Falls Laboratory, University of Minnesota. The sediment used in all the experiments was uniform
gravel, with geometric mean size Dg  7.2 mm, geometric
standard deviation Sg 1.2, median particle size D50  7.1
mm, particle size for which 90% of the sediment is ner
D90  9.6 mm, and a specic gravity of 2.55 (R  1.55).
Based on these observations they found empirical equations for the dimensionless entrainment rate E*b and particle
step length L*s as functions of the Shields parameter T*, as
follows,
E*p  0.06 (T*  0.0549)1.97

(2-90e)

L*s  44.33 (T*  0.0549)0.47

(2-90f)

and

Equation (2-90e) predicts values very close to those


observed by Fernandez Luque and van Beek (1976). On the
other hand, Eq. (2-90f) contradicts the original ideas of
Einstein (1950) who did not include a critical shear stress for
motion and assumed that the dimensionless step length L*s
100 for all ow conditions. Also of interest is the fact that the
step length is found to decrease as the ow intensity characterized by T*, increases. L*s takes values between 160 and
270 for the range of experimental conditions covered in the
experiments. Wong and Parker (2006a) argued that since the
chances of a given particle being captured and trapped into
a resting position increases with sediment transport rate, it is
reasonable to assume that the step length becomes smaller

71

when the excess dimensionless shear stress gets larger, as


indicated by Eq. (2-90f). Substituting (2-90e) and (2-90f)
into (2-90b), yields
q*  2.66 (T*  0.0549) 2

(2-90g)

This empirical t has a structure which is very similar to


several formulations presented below.
Bagnolds approach is considered next. A dimensionless bed load transport equation, such as the one implied by
Eq. (2-89), can be obtained by simply dividing both sides
of Eq. (2-81) by a characteristic length given by the particle
diameter D and a characteristic velocity given by
which yields

q


qb
D gRD



cb D b
D

ub
gRD

gRD,

(2-91)

Bagnolds (1956) hypothesis makes it possible to estimate the volumetric sediment concentration in the bedload
layer per unit bed area, given by the product of sediment
concentration cb and the thickness of the bed load layer Db,
as follows
cb Db
T* T*c

D
Md

(2-92)

where Md is a dimensionless dynamic friction coefcient


(Abott and Francis 1977; Sekine and Kikkawa 1992). Nio
and Garca (1998) have used a Lagrangian particle-saltation
model to estimate values of Md and found that it takes values
in the range between 0.25 and 0.4, which are smaller than
Md  tanF  0.63 proposed by Bagnold (1973) but are in
good agreement with laboratory observations of sand transport (Nio and Garca 1998c). Simulated values of Md for the
case of sand saltation are much closer to the corresponding
observed values than in the case of saltation of gravel (Nio
and Garca 1994). Bagnolds hypothesis used to obtain to
Eq. (2-92) would be only valid for intense transport conditions involving very high sediment concentrations of sandsize material. The mean velocity of the particles in the bedload
layer ub can also be estimated with the help of numerical
modeling and experimental observations of particle motion
(Reizes 1978; Leeder 1979; Murphy and Hooshiari 1982;
Bridge and Dominic 1984; van Rijn 1984a; Wiberg and
Smith 1985, 1989; Sekine and Kikkawa 1992; Lee and Hsu,
1994; Nio and Garca 1994, 1998b, 1998c; Lee et al. 2000,
2006; Lukerchenko et al. 2006).
Ashida and Michiue (1972) presented a macroscopic analysis that does not account for the complexity of the saltation
process, in particular the treatment of the particle collision

72

sediment transport and morphodynamics

with the bed. In their analysis, a simplied particle equation


of motion is used to obtain the following expression for the
dimensionless mean particle velocity in the bedload layer:

where the functionality is implicitly dened by the relation


*

1
ub

8.5 T*

gRD

1 2

1 2

 T*c

1 2

1 2

 T*c

(2-94)

Ashida and Michiue recommend a value for T*c of 0.05 in


their relation. It has been veried with uniform material ranging in size from 0.3 mm to 7 mm. The Ashida-Michiue bed
load transport equation is a good example of a Bagnoldean
formulation. It is very similar to the one developed independently by Engelund and Fredse (1976) and more recently by
Nio and Garca (1998) using a Lagrangian particle-saltation model for bed load transport.
In addition to the relation of Ashida and Michiue (1972),
the following bed load transport relations are of interest.
Meyer-Peter and Muller (1948):

q*  8 T*  T*c

2

(2-95a)

where T*c  0.047. This formula is empirical in nature; it has


been veried with data for uniform coarse sand and gravel. Even
though it was developed for alpine streams in Switzerland, it
enjoys wide use in coastal sediment transport (e.g. Soulsby,
1997). Recently, Wong and Parker (2006b) reanalysed the data
used by Meyer-Peter and Muller and found that a better t is
provided by one of the two alternative forms;

q* 4.93 T* 0.047

1.6

(2-95b)

t
e dt 

 0.413 T* 2

(2-93)

Upon substitution of Eqs. (2-92) and (2-93) into (2-91)


and assuming a value for Md of 0.5, the following nal form
for bed load transport is obtained
q* 17 T* T*c T*

0.413T 2

43.5 q*
1 43.5 q*

(2-96b)

This relation constitutes the rst attempt to derive a bed


load function. Note that this relation contains no critical shear
stress. It has been used for uniform sand and gravel. Gomez
and Church (1989) recommend its use for cases where the
local bed load transport rate needs to be calculated. Yang and
Wan (1991) found that it could predict sediment transport
rates in large rivers but not in small rivers and umes.
Yalin (1963):
ln 1 a2 s
q*  0.635s(T* )1 2 1

a2 s

(2-97a)

where

1 2

a2  2.45( R 1)0.4 T*c

s

T*  T*c
T*c

(2-97b)

and T*c is evaluated from the Shields curve. Two constants in


this formula have been evaluated with the aid of data quoted
by Einstein (1950), pertaining to 0.8 mm and 28.6 mm material. Wiberg and Smith (1985, 1989) were able to reproduce
Yalins relation, with their saltation-based bed load transport
model.
Wilson (1966):

q*  12 T*  T*c

32

(2-98)

where T*c is determined from the Shields diagram. This relation is empirical in nature; most of the data used to t it
pertain to very high rates of bed load transport. It has been
used extensively to estimate transport of sand and industrial
materials such as nylon in pressurized ows (e.g., Wilson
1987).
Paintal (1971):

q  3.97 T  0.0495

2

(2-95c)
q*  6.56 1018 T*16

Einstein (1950):

q*  q* T*

(2-96a)

(2-99)

was obtained though extensive measurements of very low


bed load transport rates. It is valid for 0.007 T* 0.06 and
sediment grain sizes between 1 mm (coarse sand) and 25 mm
(gravel). This relation shows that for low shear stresses, the

bed load transport


sediment transport phenomenon is highly nonlinear. That is,
for small changes in bed shear stress, the rate of bed load
transport increases dramatically.

equation is based on a dimensionless particle diameter and


the transport stage parameter T, dened, respectively, as
gR
D*  D50 2
N

Engelund and Fredse (1976):


1 2

q*  18.74 T*  T*c T*

1 2

 0.7 T*c

(2-100a)

1

 Rep

(2-103b)

and
*

where T*c 0.05.


This formula resembles that of Ashida and Michiue
because its derivation, albeit obtained independently, is
almost identical. This relation was rederived by Fredse and
Deigaard (1992, p. 214), resulting in a very similar relation,

73

T

Ts  Tc
*
Tc

(2-103c)

Here T*s is the bed shear stress due to skin or grain friction,
and T*c is the critical shear stress for motion from the Shields
diagram.
Madsen (1991):

30 *
q 
T  T*c T*

P Md
*

1 2

 0.7

1 2
T*c

(2-100b)

Fredse and Deigaard tested the formula for different


values of the dynamic friction coefcient Md. For Md  1.0,
Eq. (2-96b) gives results very close to those of the MeyerPeter and Muller formula (Eq. 2-95a). However, both formulations were found to overpredict bed load transport at hight
shear stresses.

q*  FM (T*1 2  0.7T*c1 2 )(T*  T*c )

(2-104a)

where FM  8  tan F for rolling/sliding sand grains and FM 


9.5 for saltating sand grains in water.
Nielsen (1992):
q*  12T*1 2 (T*  T*c )

(2-104b)

Fernandez-Luque and van Beek (1976):

q  5.7 T

2
 T*c

(2-101)

where T*c varies from 0.05 for 0.9-mm material to 0.058 for
3.3-mm material. The relation is empirical in nature and was
obtained through laboratory observations.

obtained by tting to uniform size sand and gravel bed load


transport data. This relation was also independently derived
by Soulsby (1997). Equations (2-104a) and (2-104b) have
been used mainly in coastal engineering.
Nio and Garca (1998b):
q* 

Parker (1979):

T  0.03
q  11.2
*

T* 3

(2-102)

(2-104c)

obtained with a Lagrangian description of bed load transport


by saltating particles and tested with experimental observations of gravel transport (Nio and Garca 1994) and sand
transport (Nio and Garca 1998c). A dynamic friction coefcient Md  0.23 was determined, almost three times smaller
than the value proposed for the same coefcient by Bagnold
(1973). This relation basically has the same structure as
Madsens Eq. (2-104a).
Cheng (2002):

Van Rijn (1984a):


T 2.1
q  0.053 0.3
D*

4.5

developed as a simplied t to the relation of Einstein


(1950) for the range of Shields numbers likely to be
encountered in gravel-bed streams. This formula was used
to analyze the hydraulic geometry of gravel-bed streams
(Parker 1979).

12
T* T* c T*12 0.7T*1c2
Md

(2-103a)

can be used to estimate bed load transport rates of particles


with mean sizes in the range between 0.2 and 2.0 mm. This

0.05
q*  13 T*  2 exp  *  2
T

(2-104d)

This relation gives results similar to those obtained with


Meyer-Peter and Muller (1948) and Einstein (1950) equation
for moderate dimensionless shear T* values. It also agrees

74

sediment transport and morphodynamics

well with the values predicted with Paintal (1971) for weak
transport conditions under low shear stresses.
For intense transport conditions associated with large
values of the dimensionless Shields stress T*, Eq. (2-104d)
reduces to
q* 13 T* 2

(2-104e)

In fact, most of the bed load transport equations display


the same asymptotic behavior for high values of shear stress
as can be observed in Fig. 2-33. That is, for T* T*c, q* z
T*32. However, a word of caution is necessary because this
does not seem to be the case according to the observations
made by several investigators who have found a different
relation for high transport rates. For instance, Rickenmann
(1991) has indicated that for intense sediment transport conditions, when T*  0.4, grain ow (or sheet ow) conditions
develop and q* is actually proportional to T*52 as found by
Hanes and Bowen (1985) with a granular uid model for
coastal sediment transport and Takahashi (1987) for debris
ows on steep slopes. Hanes (1986) showed that under these
conditions transport rates can be approximated with
q*  6 T* 2

(2-104f)

These ndings indicate a stronger dependence of q* on T*


at very high transport intensities that one might expect from
bed load transport equations. In fact, Abrahams (2003) has
recently revisited the concepts advanced by Bagnold (1973)
and found that transport rates under sheet ow (grain ow)
conditions are much higher than previously thought.
Most of the bed load equations shown above apply to
the case of mild slopes or nearly horizontal ows. For steep
channels, the effect of the dowslope gravitational component
cannot be neglected. Smart (1984), Bathurst et al. (1987),
Graf and Suszka (1987), Tsujimoto (1989), Rickenmann
(1991), Damgaard et al. (1997), and Aguirre-Pe and Fuentes
(1995), among others, have proposed equations for bed load
transport in steep slopes.
Only a few research groups have attempted complete
derivations of the bed load function in water. They include
Wiberg and Smith (1989); Sekine and Kikkawa (1992);
Nio and Garca, (1994, 1998); Seminara et al (2002); and
Parker et al (2003). These results are encouraging because
they show that bed load transport can be predicted with
a mechanics-based approach. More recently, discrete particle simulations of bed load transport which account for
the effect of near-bed turbulence, particle location and
particle-particle interaction, have been conducted, among
others, by Jiang and Haff (1993), Drake and Calantoni
(2001), Nelson et al. (2001) and Schmeeckle and Nelson
(2003). Undoubtedly, the role played by turbulence in bedload transport is still a subject that deserves more research
(Nelson et al. 1995; Garca et al. 1996; Nio and Garca
1996; Best et al., 1997; Sechet and Le Guennec 1999;
Schmeeckle et al., 2001; Papanicolau et al. 2001; Sumer
et al. 2003). It is also clear that direct eld observations
provide the best information for both developing and testing of new formulations (e.g., Almedeij and Diplas 2003).
Several bed-load transport relations for gravel and sediment mixtures are considered in Chapter 3.

2.6.5 Two-Dimensional Transport of Bed Load


The relations presented above for bed load transport are all
one-dimensional in nature. That is, they provide the magnitude of a bed load transport vector that is oriented in the
direction of the boundary shear stress. That is, if the s coordinate is directed along the bed parallel to the boundary
shear stress and the n coordinate is directed along the bed
and perpendicular to the s coordinate,
G
q  (qs , qn )  (q ,0)

Fig. 2-33.
literature.

Plot of several bed load functions found in the

(2-105)

G
where q denotes the two-dimensional vector of bed load
transport rate and q denotes the magnitude of that vector,
which is computed using one of the relations presented
above.

You might also like