You are on page 1of 23

M O D E L FOR CALCULATING B E D L O A D

Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

TRANSPORT O F SEDIMENT
By Patricia L. Wiberg 1 and J. Dungan Smith2

ABSTRACT: A bed load transport model based on the mechanics of sediment mov-
ing by saltation yields predicted values of bed load flux as a function of boundary
shear stress, grain diameter and density. The parameters required to calculate bed
load transport (particle velocity, bed load sediment concentration, and the height
of the bed load layer) can all be determined from our model, which computes
sequences of trajectories of individual saltating grains as well as the concentration
of moving grains that the flow can support. The latter is related to the momentum
the accelerating grains extract from the flow. Predicted curves of bed load transport
vs. boundary shear stress agree well with transport data collected by Gilbert, Meyer-
Peter et al., and the Waterways Experimental Station; measured shear stress was
corrected for pressure drag when bed forms were present. Comparison of predicted
bed load transport with common bed load equations reveals considerable similarity
among the relationships. The best agreement with the data is produced by rela-
tionships in which the transport rate vanishes as the shear stress approaches the
critical Shields' number, as in our model and Yalin's equation.

INTRODUCTION

Most previous research on bed load transport of sediment has focused on


developing empirical or semi-empirical expressions for estimating transport
rates in unidirectional flow. Empirical expressions such as the Meyer-Peter
and Miiller equation (1948) typically have been obtained by using dimen-
sional analysis or physical reasoning to produce nondimensional variables
that collapse experimental transport data into a simple relationship. The semi-
empirical bed load equations, for example those of Einstein (1942, 1950)
and Yalin (1963), derive their structure, in part, from considerations of the
mechanics of sediment moving in the saltation mode. While the latter ap-
proach has been valuable for formulating bed load equations and for ex-
amining the mechanics of saltating particles, the semi-empirical expressions
include one or more coefficients that have to be set using bed load transport
measurements. Consequently, neither of these types of equations is general
enough to permit reliable, predictive calculations of transport for sediment
types or flow conditions that are substantially different from those studied
experimentally.
The bed load transport model presented herein approaches the problem
from a mechanical point of view, but yields transport rate predictions without
reference to empirical bed load transport data. The foundation for our ap-
proach is a saltation model that determines the paths of single grains moving
as bed load; it is described by Wiberg and Smith (1985). The heights and
lengths of the trajectories made by hopping grains, as well as the grain ve-
locities along these trajectories, can be determined from the saltation model
'Res. Assoc., Dept. of Geological Sciences, AJ-20, University of Washington,
Seattle, WA 98195.
2
Professor, Geophysics Program, AK-50, University of Washington.
Note. Discussion open until June 1, 1989. To extend the closing date one month,
a written request must be filed with the ASCE Manager of Journals. The manuscript
for this paper was submitted for review and possible publication on February 20,
1988. This paper is part of the Journal of Hydraulic Engineering, Vol. 115, No.
1, January, 1989. ©ASCE, ISSN 0733-9429/89/0001-0101/$1.00 + $.15 per page.
Paper No. 23123.

101

J. Hydraul. Eng., 1989, 115(1): 101-123


and the results have been shown to agree well with available measurements
(Francis 1973, Abbott and Francis 1977). Quantitative information about
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

particle trajectories, however, is not very extensive and the ultimate test of
any saltation model comes from using it to predict bed load transport rates
for which considerably more data are available. The flux of bed load sedi-
ment is given by the product of the downstream particle velocity and the
sediment concentration integrated over the height of the bed load layer. Ex-
panding the saltation model to produce a bed load transport model thus re-
quires a method for calculating the concentration; this is described in the
first section of the paper. All other parameters needed to calculate bed load
transport rates are available directly from the saltation model. In the follow-
ing sections, we compare the results of our model with existing flume mea-
surements of bed load transport rates, as well as with several commonly used
bed load transport equations.

THEORETICAL CALCULATIONS OF B E D LOAD TRANSPORT

The volume discharge of bed load sediment per unit width of flow, qs, is
given by the equation,
/•(8«) m

qs = csu4z (1)
J-n
where cs = cs(z) is the concentration of sediment in the bed load layer, z is
height above the bed, us = us{z) is the downstream (bed-parallel) component
of sediment velocity, -n is the height of the nonmoving bed surface, and
(8B),„ is the height of the bed load layer (maximum saltation height). The
saltation model described by Wiberg and Smith (1985) yields the trajectories
of single grains over a sediment bed by solving particle equations of motion
in which the acceleration of the grain is balanced by the gravitational and
fluid mechanical forces acting on it (lift, drag, and relative acceleration).
The governing equation, in vector form, is
d(us) d(Vf) CVAD . 3<uA)
psV —— = pV —-— + p —— uA uA + pVC,„ ——
dt dt 2 at
C A
+ 9-y1 l(ul)T- (ul)B]eL - (p, - p)VQ (2)

where uA is the difference between the fluid and particle velocity, uf — us,
( ) indicates an average over the grain, and the subscripts T and B refer to
the top and bottom of the grain; V is particle volume, AD and AL are particle
cross-sectional area normal to the drag and lift forces, respectively, p and
ps are fluid and sediment density, g is gravitational acceleration, and eL is
a unit vector in a bed-normal direction. The drag coefficient, CD, and the
coefficient of relative acceleration (virtual mass), C,„, are taken as those for
spheres, and the lift coefficient CL is set equal to 0.2 (see Wiberg and Smith
1985). In the derivation of Eq. 2, both RD =? |uA|D/v and R = ufh/v have
been assumed to be significantly greater than one; here, D is the nominal
sediment diameter, v is the kinematic viscosity of the fluid and h is the flow
depth. Neither of these constraints pose significant restrictions on the in-
tended use of Eq. 2. (For details, see Wiberg and Smith 1985.)

102

J. Hydraul. Eng., 1989, 115(1): 101-123


Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

4r
3-

D *•

0
20 40 60 80 100 120 140
X/D

FIG. 1. (a) Typical Calculated Trajectory, with Magnitudes of Forces at Several


Locations Indicated by Length of Arrows (2.1 Vertical Exaggeration); (b) Typical
Hop Sequence

The solution to these equations gives the position of a saltating particle


as a function of time, from which trajectory heights, lengths, and particle
velocities can be determined. A typical, calculated trajectory is shown in
Fig. 1(a). The relative magnitudes of the forces acting on the particle at
several points along the trajectory are indicated by the lengths of the arrows;
FL is the lift force, Fg> is the gravitational force, FD is the drag force, and
Fv is the force due to acceleration differences between the fluid and the
sediment; FD and Fv have both horizontal and vertical components. The ini-
tial conditions for the first hop a particle makes after resting on the bed are
calculated from the fluid-mechanical forces acting on it. During subsequent
collisions with the bed, however, the particle retains some of its momentum,
bounces back into the flow, and continues to move in a succession of hops
until the lift-off angle or velocity is such that it cannot continue. The col-
lision between the moving grain and a grain on the bed surface can occur
at any point on the exposed, upstream side of the stationary grain. The model
allows the point of impact to vary randomly, thereby adding a stochastic
element to the solution. A typical sequence of hops calculated with the sal-
tation model is shown in Fig. 1(b).
The time a particle spends at any level above the bed is related to the
probability of it being at that level. This in turn is one way of defining the
concentration of sediment as a function of height above the bed, i.e., cs(z)
= (dt/dz)D, where the subscript indicates that at any given time the particle
extends over a finite vertical distance D. Under uniform, steady flow con-
ditions, profiles of (dt/dz)D for a large number of hops (on the order of 100)
can be summed to give a vertical profile of sediment concentration above a
point on the bed. Normalizing the resulting profile so that the integral of

103

J. Hydraul. Eng., 1989, 115(1): 101-123


the concentration over the height of the bed load layer is unity yields the
vertical structure of the concentration profile, i.e., c*(z) = y2,'l=r(dt/dz)D.
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

While the absolute magnitude of the concentration is not known yet, the
sediment concentration in the bed load layer can be written as cs = (cs)c*,
where (cs) is the vertically averaged value of the concentration.
In order to evaluate the magnitude of the sediment concentration, it is
necessary to consider what limits the number of grains that can be trans-
ported in the bed load layer. When a particle leaves the bed and is accel-
erated by the downstream component of the fluid flow, it extracts momentum
from the fluid, locally reducing the shear stress of the fluid. At some con-
centration, the sediment will extract enough momentum near the bed to re-
duce the boundary shear stress of the flow to the critical value and no ad-
ditional net erosion of particles will occur. If enough of the saltating grains
lodge in the bed to allow the local boundary shear stress to rise above the
critical value, additional particles will be mobilized until the boundary shear
stress again decreases back to the appropriate value. Owen (1964) pointed
out the importance of this self-equilibrating process and equated the critical
shear stress for this process to that for initial motion. He then used this
condition to provide a functional relationship between sediment concentra-
tion and boundary shear stress in a study of eolian saltation.
The first step in applying Owen's approach to the problem of determining
sediment concentration is to partition the total stress in the bed load layer
into the stress the fluid exerts on other fluid parcels, iy, and the stress the
sediment exerts on the fluid, T„ i.e.:
Ttotal = T/ + Ts (3)

If the thickness of the saltation layer is much less than the flow depth, the
total stress in this thin layer above the bed is approximately constant and
equal to the shear stress of the fluid, ib, just above the layer of saltating
grains. The sediment component of stress depends on the amount of mo-
mentum transferred from the fluid to sediment during each trajectory and
the concentration of saltating grains. The transfer of momentum from the
fluid to the sediment, resulting in downstream particle acceleration, is pri-
marily accomplished by the horizontal component of the drag force, FD =
(p/2)CDAD\uA\uA. This drag force component has a minimum value near the
top of the bed load layer where the sediment velocity approaches the fluid
velocity, and is maximum near the bed at the beginning of the trajectory
where the difference between the particle and fluid velocities is the greatest.
The sediment stress is equated to the downstream component of drag force
per unit area, where the appropriate area is the grain diameter times the
average downstream distance between two grains. This distance can be found
from the definition of sediment concentration as the ratio of sediment volume
to total volume (fluid plus sediment). Taking the sediment volume equal to
the volume of a single grain yields the average fluid volume between two
grains, V/cs. The grains can be anywhere in the volume, so two grains will
not, in general, lie in the same cross-stream plane or be at the same level
above the bed. If we wish to calculate the average distance between two
grains "lined-up" in the downstream direction, then this volume must be
divided by D2, yielding a distance of /(z) = V/[D2cs(z)], an area of ID, and
a stress of

104

J. Hydraul. Eng., 1989, 115(1): 101-123


FD
FD{z) AD FDaD
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

l(z)D V AD
DAD
The geometric factor aD = ADD/V in this equation is 1.5 for spheres. Eval-
uating Eq. 3 at the top of the nonmoving grains of the bed surface, -n, gives
ftotai — T 6 = T>(T|) + T S (T|). If the fluid stress at the bed is equal to the critical
shear stress, as argued above, the sediment stress at the bed can be written
as T/TI) = T 6 — Tcr. Substituting Eq. 4 for T, produces an equation in which
everything is known except the vertically averaged sediment concentration.
Solving for this yields,

Jb
<c.) = ~ ] \ (5)
c*(iq) :
AD
All of the terms in Eq. 1 necessary to calculate bed load sediment flux
can now be evaluated using the saltation model. Figs. 2(a)-(c) show the
average saltation height, ( 8 ^ ) ^ / 1 ) , the average downstream particle velocity,
(Wj)avg/"*> a n < i m e vertically averaged bed load concentration, (c s ), as a func-
tion of transport stage, T* = T 4 /T„., for five sediment sizes that will be the
focus of the next section: D = 0.035 cm, 0.050 cm, 0.080 cm, 0.33 cm,
and 2.86 cm. The sediment density is taken to be ps = 2.65 g/cm 3 in all
cases. Fig. 2(d) shows the calculated volume flux of bed load sediment, qs
(in units of cm 3 /cm-j), as a function of transport stage for the same five
cases.

COMPARISON WITH B E D LOAD TRANSPORT MEASUREMENTS

Verification of the bed load transport model described above must be made
by comparing bed load flux calculations to carefully made measurements of
bed load transport rates obtained under appropriate flow conditions. In par-
ticular, the flow should be steady, two-dimensional, and horizontally uni-
form. When ripples are present on the bed, the upper part of the flow should
be uniform with the bed roughness scaled by the bed form size; the inner
flow will not be locally uniform, but the spatial average of the inner-flow-
induced boundary shear stress can be related to the sediment transport rate
(Smith and McLean 1977). The parameters required for the model are grain
size and density, water temperature and density, and boundary shear stress.
Most of the sediment transport data recorded under suitable flow conditions
comes from flume studies; for the present purpose, the most useful of these
were conducted by Gilbert (1914), Meyer-Peter et al. [(1934); the Meyer-
Peter et al. data employed in this paper were taken from Brownlie (1981)]
and the U.S. Waterways Experimental Station (1935). There are many other
bed load transport studies in the literature, but for the most part they are
limited in scope and add little to these three data sets, or are inappropriate
for some other reason. For example, Guy et aL (1966) conducted an exten-
sive study that focused primarily on bed form evolution, but it is, in large
measure, unsuitable for the present purpose because of the three-dimensional

105

J. Hydraul. Eng., 1989, 115(1): 101-123


o D = 0.035 cm
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

FIG. 2. Bed Load Transport Parameters As Function of Transport Stage, r*, for
Five Sizes of Quartz-Density Sediment: (a) Mean Saltation Height; (b) Average
Downstream Particle Velocity; (c) Vertically Averaged Concentration; (d) Volume
Flux of Sediment

nature of the flow. One study that does add to these comprehensive inves-
tigations is that of Fernandez Luque and van Beek (1976), who examined
lower plane-bed transport for which little other data is available.
A number of sets of measurements have been chosen from the Gilbert,
Meyer-Peter et al., U.S. Waterways Experimental Station (WES) and Fer-
nandez Luque and van Beek data sets for the purpose of testing the sediment
flux calculations described above. The selection was made to include a wide
range of sediment sizes with each size represented by a reasonable number
of bed load transport measurements covering as large a range of shear stresses
as possible. In some cases this was accomplished by merging measurements
for the same sediment size made by several investigators; good agreement
106

J. Hydraul. Eng., 1989, 115(1): 101-123


in the overlapping portions of the data sets was used to confirm that the data
were free of significant systematic errors. In combining sets of measure-
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

ments made either by one or several investigators, grain sizes that were roughly
the same, i.e., within about 10% of each other, were grouped together pro-
vided that the differences in their nondimensional critical shear stress values,
(T*)cr = T„/[(ps — p)gD], were also less than 10%.
Nondimensional sediment transport cj> = qs/[(ps/p — l)gD 3 ] 1/2 is plotted
as a function of nondimensional shear stress, T* = ib/[{ps — p)gD] in Fig.
3(a) for the combined data of WES and Gilbert for D = 0.05 cm, p, = 2.65
g/cm 3 . These measurements cover the whole range of bed load transport
from initial motion to upper plane-bed transport. The data fall into two dis-
tinct ranges of shear stress for relatively low transport rates. The lower-left
set of points corresponds to sediment transported in the lower plane-bed phase
(over a flat bed). As the shear stress on the bed increases, the transport rate
increases until a transport stage T* = T6/Tcr of 2 to 2.5 is reached, at which
point ripples begin to form. The ripples produce a drag on the flow, denoted
here as form drag, that reduces the shear stress on the true bed, denoted
here as skin friction. It is the skin friction, T^, that scales the forces on the
sediment grains in the neighborhood of the bed, and hence, that is related
to the sediment transport rate. While the shear stress of the outer flow, T„,
increases as the ripples form, the stress available to transport sediment de-
creases and there is a commensurate drop in the transport rate. Once ripples
cover the bed, the transport rate again increases with an increase in T„; the
lower-right set of points in Fig. 3(a) correspond to measurements made on
a rippled bed. Skin friction can be determined with reasonable accuracy from
the outer shear stress over rippled beds provided the bed forms are two-
dimensional and their heights and wavelengths are known; the procedure will
be described briefly in the discussion of the WES data below.
In the upper portion of the plot, the curve formed by the bed form data
appears to merge with an extension of the lower plane-bed data. Here, the
bed is essentially planar again (upper plane-bed) and form drag is no longer
important. This transition to upper plane-bed conditions can occur in several
ways depending on the flow conditions. First, as the shear velocity begins
to approach the settling velocity of the saltating particles, turbulent velocity
fluctuations start to affect the particle trajectories, making them significantly
longer. Under these conditions, denoted incipient suspension, the relatively
short wavelength bed forms begin to decay in amplitude and the form drag
on these features is reduced, thereby automatically increasing the effective
shear stress on the bed and causing an increase in the intensity of turbulence
on the upper part of the ripples. Thus, once the bed forms begin to decay
from their equilibrium forms, the transition to a flat bed proceeds rapidly.
The bed can also go through the transition from rippled to flat bed because
of limited flow depth, even if the conditions for incipient suspension have
not been reached. The maximum amplitude of two-dimensional bed forms
is restricted roughly to less than one-fifth of the depth of flow. In shallow
flows where the bed forms quickly reach their maximum heights, increases
in transport stage result in longer wavelengths and thereby can cause the bed
forms eventually to become so long compared to their amplitude that they
are essentially indiscernible. The maintenance of the bed forms by the flow
is also dependent on Froude number, F = (w)/Vgft, which varies inversely
with depth, h. As the Froude number increases toward the critical value F

107

J. Hydraul. Eng., 1989, 115(1): 101-123


Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

10' I01 : e

• D = 0.05lcm (Gilbert) : upper plane - bed M

• o D = 0.05lcm,0.052cm(WES)
10° I0C r
/i

o 6° /oo
JO" 1 r 10" r
of °
$
„ £o with bedform
10" io- - o o o „2,o> correction
o o 3»j
03
: o
0
o /<B

10" 10" r O Jo <D


eoo

J. Hydraul. Eng., 1989, 115(1): 101-123


• b.
. . I 1 • • • • . . , , i
10-* i o - 4 2, z
10-2 4 6 8 |Q -1 2 4 6 8 |Q0 io- 4 6 8 |Q -I 2 4 6 6 |0°

F!G. 3. Nondimensional Bed Load Transport, 4>, as Function of Nondimensional Shear Stress, T*, f o r D = 0.05 cm: (a) Data with Calculated
Curve; (b) Data after Bed Form Correction.
= 1, the factors controlling bed stability are altered and the bed may become
essentially planar before antidunes develop at supercritical Froude numbers.
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

The latter two situations are particularly likely to exist in flume channels
where the flume dimensions, pump capacities, and sidewall effects limit the
flow depth that can be used. The decrease in the ratio of bed form height
to wavelength in either of these shallow-flow situations again results in re-
duced form drag and increased shear stress at the bed; however, sediment
can still be moving predominantly in the saltation mode.
Gilbert (1914), Meyer-Peter et al. (1934), and the Waterways Experi-
mental Station (WES) personnel (1935) measured transport rates of sand and
gravel in parallel-sided flumes. Combinations of water depth, surface slope,
flow discharge and flume width were employed to produce a variety of shear
stresses in the experiments. The boundary shear stress of a flow over a flat
bed is given by the product of the flow depth and surface slope, ds/dx, i.e.,
T(, = — pgh(ds/dx). While the hydraulic radius (wetted perimeter divided by
the cross-sectional area of flow) is often used instead of depth when cal-
culating shear stress in natural channels where the roughness of the sides is
comparable to the bottom, the depth is a more appropriate parameter for
flumes with glass sides because the walls are much smoother hydrody-
namically than the bottom. If the flume width is not sufficiently large with
respect to the depth, however, the effect of friction on the sidewalk can
significantly affect the surface slope. This effect of sidewall friction can be
removed from flume data using a sidewall correction such as Williams' (1970)
experimentally determined adjustment,

^measured
^corrected ~ ~. T~> y^)

where b is the flume width; the coefficient 5.5 is consistent with length
measured in centimeters (0.055 in SI units). In the analysis that follows, Eq.
6 was employed to correct all shear stress measurements, but data requiring
more than a 5% correction were discarded to minimize error associated with
uncertainties in adjusting for sidewall effects.
Gilbert (1914) measured transport rates for eight sizes of sand and gravel
with diameters ranging from D = 0.03 c m t o f l = 0.49 cm. These mea-
surements cover a particularly large range of shear stresses, in many cases
including upper plane-bed transport and antidunes. Unfortunately no mea-
surements were made of the height and wavelength of bed forms, so only
the lower and upper plane-bed measurements can be related to the skin fric-
tion on the bed. Many of the flows were quite shallow (~5 cm), so the
ripples and dunes were almost certainly depth limited. For the most part,
the upper plane-bed data correspond to supercritical flow conditions, and
with the possible exception of the D = 0.03 cm case, incipient suspension
was not responsible for the observed transition from dunes to upper plane-
bed, which occurs at a roughly constant transport stage of nine to ten. It
appears, therefore, that the sediment was transported primarily by saltation
in Gilbert's upper plane-bed runs. The grain sizes in the runs used for the
model comparison are D = 0.03 cm, 0.05 cm, 0.08 cm, and 0.32 cm [grades
A, C, D, and F in Gilbert's (1914) notation].

109

J. Hydraul. Eng., 1989, 115(1): 101-123


The Waterways Experimental Station (1935) made nine series of sediment
transport measurements in a 70 c m wide, tilting flume using sand and gravel
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

ranging in size from D = 0.03 c m to 0.41 cm and one set of measurements


over a smooth cement bottom. The shear stress was varied from below the
critical value well up into the ripple and dune range, but never reached upper
plane-bed values. Streamwise profiles of the rippled beds were made pe-
riodically for all but one measurement series and the data for five series
include information on surface velocity as well as mean velocity and water
temperature. This makes the WES data particularly valuable for comparison
with the bed load flux calculations; their D = 0.05 cm and 0.03 cm diameter
sets of transport data are used in the present comparison (sands 3, 4 , 6, 7
in WES notation). The measurements made on rippled beds, however, have
to be corrected for form drag before they can be compared to the model.
The flow over a rippled bed can be decomposed into two parts. The outer
flow extends from the surface down to some height above the bed, z* (the
matching height), and the inner flow extends from z* to the bed. The level
z* is related to the average height of the internal boundary layer that forms
in the wake of the upstream bed form. The shear stress for the outer flow
is given by T„ = — pgh(ds/dx), and the corresponding roughness parameter,
(z0)„, is a function of the heights and wavelengths of the bed forms. The
spatially averaged velocity profile for the inner flow can also be approxi-
mated by a logarithmic profile (Smith and McLean 1977), but in this case
z0 depends on the local roughness of the bed, and the shear stress is equal
to the shear stress at the bed, i.e., the stress of the outer flow minus the
form drag associated with the flow over the bed forms. The drag force can
be written as FD = (p/2)CDu2rHb, where CD is now the drag coefficient for
the bed form, ur is the reference velocity of the undisturbed flow, and H is
bed form height. Taking the reference velocity to be the average of the un-
disturbed flow velocity over the vertical extent of the bed form gives ur =
(«*), / K _1 {ln[^/(z 0 ) J /] - 1}; recalling that the skin friction TS/ = tb = p(w*)?/,
yields

FD f CDH\ H ]V'

where \ is the bed form wavelength. The drag coefficient over bed forms is
not well known, but velocity measurements over large dunes by Smith and
McLean (1977) indicate that when separation occurs just downstream of the
bedform crest, the drag coefficient CD is approximately 0.2. The roughness
parameter of the inner flow, (z0)sf, depends on the local roughness length
scale of the bed. Applying this correction to the WES data generally yields
good results. Fig. 3(b) shows the same WES data plotted in Fig. 3(a) after
the bed form correction is made. The correction brings the bed form data
into good agreement with the lower plane-bed data, and at higher transport
stages the data trend up toward the upper plane-bed values.
The Meyer-Peter et al. (1934), see Brownlie (1981), data for sand and
gravel include measurements for a number of grain sizes over a relatively
narrow range of shear stresses, in some cases primarily in the bed form
range; no information about bed form size is given. The data set is notable
in that it was, in part, the basis of the well-known Meyer-Peter and Miiller
(1948) and Einstein bed load equations. The 2.86 cm data of Meyer-Peter
110

J. Hydraul. Eng., 1989, 115(1): 101-123


Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

I0 1 I0 1
• • D = 0.031cm, 0.037cm upper plane-bed M%%
(Gilbert)
l 0 o L O D = 0.03lcm,0.035cm
10° r
(WES)

io- io-'-

with bedform
io- 10" ° SP
of correction

10" 10"
Ioo o oo O OCfflO O

J. Hydraul. Eng., 1989, 115(1): 101-123


_ l I 1 _
10- lO" 4
|Q-2 2 4 6 8 IQ-I 2 4 6 8 |Q0 IQ-2 2 4 6 8 |0-l 2 4 6 8 |Q0

FIG. 4. Nondimensional Bed Load Transport as Function of Mondimensional Shear Stress for I) = 0.035 cm: (a) Data with Calculated
Curve; (b) Data after Bed Form Correction.
10'
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

• D= 0.079cm (Gilbert)
* D = 0.09cm (Fernandez Luque)
C
I0

10" I _

4>
10- 2 _

10 - 3

10"
io- 6 8 10 -I 6 8 I0C

FIG. 5. Data and Calculated Curve of Nondimensional Bed Load Transport as


Function of Nondimensional Shear Stress for D = 0.08 cm.

et al., which is all lower plane-bed, will be used in the subsequent com-
parisons. More recently, Fernandez Luque and van Beek (1976) measured
bed load transport rates for coarse sand and gravel at low transport stages.
All of their measurements were made on a planar bed, although in many
cases the bed was tilted between 12° and 20°. Their data for D = 0.09 cm
and 0.33 cm with a bed slope of zero are used to provide additional low
transport rate data for comparison with the bed load model; each of their
values represents an average of at least 20 measurements.
Figs. 3-6 show the combined data of Gilbert, WES, Meyer-Peter et al.,
and Fernandez Luque and van Beek for grain sizes, D = 0.035 cm, 0.05
cm, 0.08 cm, 0.33 cm and 2.86 cm, with density ps = 2.65 g/cm 3 . The
rippled-bed data for 0.035 cm and 0.05 cm sizes, corrected for form drag,
are plotted in Figs. 3(b) and 4(b); the Gilbert bed form data is not included
in these plots because insufficient information is available to make the form-
drag correction. Gilbert's shear stress values forD = 0.08 cm and 0.32 cm
when bed forms were present, which again could not be corrected, are in-
dicated in Figs. 5 and 6(a) by the small symbols. In each figure the solid
112

J. Hydraul. Eng., 1989, 115(1): 101-123


Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

10' 10'
• D=0.32cm(Gilbert) x D = 2.86 cm (Meyer-Peter et al.)
* D= 0.33cm(Fernandez Luque)/
10° 10°

10" 10" I _

10-2 10"

io- 10- 3

J. Hydraul. Eng., 1989, 115(1): 101-123


. . . .1 lO" 4
IQ-I |Q-2 |QO
10" IQ-2 2 4 6 8 2 4 6 6 |gO 2 4 6 8 |Q-I 2 4 6 8

FIG. 6. Data and Caiculated Curve of Mondimensional Bed Load Transport as Function of Nondimensional Shear Stress for: (a) D = 0.33
cm; (h) D = 2.86 cm.
curve shows the calculated relationship between the nondimensional volume
transport, <)> = qs/[(ps/p - l)gD 3 ] 1/2 , and nondimensional shear stress, T#
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

= T(,/[ps - p)gD].
In general, the calculated curves are in good agreement with the lower
plane-bed data, and extend upward into the upper plane-bed data. The ap-
parently planar bed values that lie to the right of the curve for D = 0.33
cm border on the bed form values and may have included subtle bed form
features. The calculated transport rates for the upper plane-bed, D = 0.035
case appear to overpredict the measured transport. This might be associated
with two effects not included as yet in the bed load model. First, the number
of particles moving in the bed load layer becomes quite large at the higher
transport stages [Fig. 2(c)] and interparticle collisions may become impor-
tant. Secondly, the extraction of momentum from the flow by the sediment
modifies the velocity profile in the bed load layer to some extent (see the
previous discussion of calculating the concentration). It is not obvious what
effect the intergranular collisions would have on the transport rate; however,
the extraction of momentum would tend to decrease the drag force on the
grains, thereby decreasing the particle velocity and, perhaps, the flux rates.
It is likely that the latter effect significantly outweighs the former, and that
the calculated fluxes for saltating sediment can be considered an upper bound
on the actual flux rates.

ANALYSIS

The good agreement between the calculated transport curves and the ex-
perimental data illustrated in Figs. 3-6 lends strong support to the validity
of the saltation/bed load transport model described in the first section. As
mentioned previously, Einstein and Meyer-Peter and Miiller used portions
of the data represented in these figures to set the coefficients in their well-
known bed load equations. It is interesting to compare these and several
other expressions for bed load transport to the calculated curves. Fig. 7 shows
a number of bed load transport expressions plotted with the corrected D =
0.05 cm data [from Fig. 3(b)] and the calculated curve. Meyer-Peter and
Miiller (1948) found that the relatively simple relationship equivalent to $
= 8(T* - 0.047)3/2 fit the Meyer-Peter et al. (1934) D = 0.52 cm and D
= 2.86 cm gravel data and Gilbert's (1914) gravel data (D = 0.32 cm, D
= 0.49 cm and D = 0.70 cm). Fernandez Luque and van Beek (1976) used
a similar equation, <> j = 5.7[T* - (T*)„.] 3/2 , to fit their data for low transport
rates (7% < 2.5). Aside from the lower value of the coefficient, the Fer-
nandez Luque and van Beek equation differs from that of Meyer-Peter and
Miiller in that the nondimensional critical shear stress is explicitly included;
the constant 0.047 in the latter equation is essentially an average value for
the nondimensional critical shear stress over the range of sizes employed in
the experiments.
Einstein (1942, 1950) formulated his semi-empirical transport expression
by relating the rate of bed load transport to the number of particles passing
through a unit cross section normal to the flow at the bed. The flux of par-
ticles was expressed in terms of the probability that a grain will experience
sufficient lift to leave the bed and that it will complete a certain number of
hops before it stops moving. This yields a functional relationship between
nondimensional transport, 4>, and nondimensional shear stress, T* with two
114

J. Hydraul. Eng., 1989, 115(1): 101-123


10'
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

10°

IO-1

<p

io- 2

IO"3

lw
|Q-2 2 4 6 8 |Q-I 2 4 6 8 |Q0

FIG. 7. Comparison of Several Weil-Known Bed Load Transport Equations with


Calculated Curve and Data for D = 0.05 cm.

unknown coefficients, A*4>*/(1 + A**))*) = /(Z?*/T*) (Einstein 1950); for


uniform sediment, T* = T* and ()>* = (|>. The constants A* and B* were set
using the Meyer-Peter et al. (1934) D = 2.86 cm and Gilbert (1914) D =
0.08 cm data. The Einstein equation is most conveniently used in graphical
form (Einstein 1950, Graf 1971).
Yalin (1963) also considered the mechanics of the transported sediment
in developing his bed load equation. He approached the problem by setting
the transport rate, q„ equal to the product of the concentration of material
moving over a unit area of the bed surface times the downstream velocity
of the material. The sediment concentration was taken to be proportional to
the excess shear stress, S = (jb — T„)/Tcr, and the velocity was estimated
from the equations of motion for a saltating particle. [See Yalin (1963) for
a description of his saltation model.] Yalin's analysis produced the equation
4> = a , 5 V ^ [ l - (flaS)"1 In (1 + a2S)], where a2 = 2.45(p/p s )°VCrX
The constant ax = 0.635 was set with the data used by Einstein in his 1942
paper, which included the D = 0.08 cm and D = 2.86 cm data shown in
Figs. 5 and 6(b).
All of these expressions plot inside the relatively large scatter of the lower
transport-rate data in Fig. 7. The Fernandez Luque and van Beek rela-
115

J. Hydraul. Eng., 1989, 115(1): 101-123


TABLE 1. Comparison of Four Bed Load Equations with IVIodei and Data for Five
Siies of Quartz Density Sediment
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

D Einstein Meyer-Peter Fernandez Luque


(cm) Model Yalin (1950) and Muller and van Beek
d) (2) (3) (4) (5) (6)
(a) T* Calculated by Method Closest to Original Analyses
0.035 0.08" 0.12 0.09 0.12 0.11
0.05 0.07* 0.09 0.09 0.13 0.11
0.08 0.05" 0.06 0.08 0.13 0.12
0.33 0.04 0.09 0.03* 0.03* 0.07
2.86 0.06 0.12 0.04 0.03* 0.12
(b) T* Calculated as -pghds/Sx
0.035 0.08 0.06* 0.10 0.10 0.11
0.05 0.07* 0.08 0.09 0.09 0.11
0.08 0.05* 0.07 0.07 0.06 0.12
0.33 0.04 0.02* 0.07 0.09 0.07
2.86 0.06" 0.06* 0.18 0.18 0.12
*Best agreement for grain size.

tionship underestimates the high transport data, but this is not surprising
since the constant in their equation was set for lower plane-bed transport
data. The constant in the Meyer-Peter and Muller equation was set by com-
parison to higher transport rates, and yielded a higher value. Pursuing this
one step further, we note that Wilson (1966) made some transport measure-
ments for upper plane-bed conditions only and fit a Meyer-Peter and Miiller-
type equation to it obtaining a still larger coefficient, § = 12[t* - (T*)cr]3/2.
The variation of the constant for these three experiments suggests that the
coefficient relating 4> and [T# — (T*) C J 3 / 2 is a function that increases with
shear stress, and that a modified version of the Meyer-Peter and Muller equa-
tion, 4> = 7[T* — (T*)CJ 3/2 . where 7 is a function of T# — (T*)cr that varies
roughly between 5 and 15, would be likely to produce good results. It is
also noteworthy that all of these empirical or semi-empirical bed load trans-
port equations fall in the range of the corrected bed form data, so that, at
least implicitly, these expressions yield transport rate as a function of skin
friction rather than the outer shear stress.
Table 1 compares the agreement between the data and the various curves.
To do this, we calculated the root-mean-square difference between the mea-
sured values of T* and the values of T* computed from the various transport
expressions for each measured value of <j>, i.e., {2[log(T*)meos - log(T*)cafc]2/
N}1'2, where N is the number of measurements. Although boundary shear
stress is the independent variable in bed load equations, the largest error in
the measurement is associated with the shear stress rather than the sediment
transport rate. In all of the cited measurements, boundary shear stress is
determined indirectly in terms of surface slope, depth or hydraulic radius,
and flume width. The D = 2.86 cm data displayed in Fig. 6(b) clearly il-
lustrates the uncertainty in the value of T*. Thus, for this comparison of the
equations to the data, the difference is computed in terms of the boundary
shear stress. Corresponding values were calculated in terms of the difference
116

J. Hydraul. Eng., 1989, 115(1): 101-123


in the transport rate (f> for the sake of completeness and resulted in virtually
the same pattern of agreement as indicated in Table 1.
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

Aside from the measurement error, the use of depth, h, or hydraulic ra-
dius, R, to calculate boundary shear stress can produce significant differ-
ences in the computed stress if the channel width to depth ratio is not large
enough that R = h. In addition, differences in the way each investigator
corrected for side-wall effects and bed form drag also affect the computed
shear stresses in the measurements. The values of the shear stress represented
in Table 1(a) are computed in the manner that most closely agrees with each
investigator's original analysis. For the model values, shear stress was com-
puted in terms of depth, with Williams' sidewall correction and the bed form
correction given by Eq. 7. Einstein (1950) and Meyer-Peter and Miiller (1948)
both used a side-wall-corrected value of R and a bed-roughness adjustment
to calculate shear stresses for the data they used to determine the coefficients
in their respective equations. The corrections used by Einstein and by Meyer-
Peter and Miiller were different, but produce almost identical values of T*.
Yalin (1963) used depth in the derivation of his equation, but used Einstein's
1942 data, in terms of hydraulic radius, to set his free coefficient; shear
stresses computed from the data by Einstein's method were compared to
Yalin's equation to produce the variances shown for Yalin in Table 1(a).
Fernandez Luque and van Beek (1976) apparently used a corrected hydraulic
radius in the analysis of their plane-bed transport data; the overall variance
is the same whether depth or hydraulic radius is used, the former being
employed to produce the values in Table 1(a). Only plane-bed and bed form
or bed roughness corrected data were used to calculate the variances in Ta-
ble 1.
Table 1(b) compares the agreement between the data and the equations
when boundary shear stress is calculated as T = — pghds/dx in all cases. As
argued previously, shear stress calculated in terms of depth will give the
best estimate of the actual boundary shear stress for channels in which the
sides are significantly smoother than the bottom, as in a glass-walled flume
or a natural channel with bed forms, although it may still be necessary to
make a side-wall correction. In all cases in Table 1(b), the transport rela-
tionships predicted by Yalin and by our model result in the smallest differ-
ence between the data and the curves. This is largely because the transport
rate goes to zero in these relationships as jb approaches the critical shear
stress for initial motion, Tcr, as is observed in the data when shear stress is
computed in the manner described here. In contrast, both the Einstein and
the Meyer-Peter and Miiller equations are invariant when plotted on 4> — T#
axes. The transport rate also goes to zero as T6 approaches J„ in the Fer-
nandez Luque and van Beek expression, but as discussed above, their re-
lationship was based on low-transport-rate data and significantly underesti-
mates the observed transport at higher transport stages. A modified version
of the Meyer-Peter and Miiller equation including the critical shear stress
and a coefficient that is a function of shear stress could be expected to work
well.
The Meyer-Peter and Miiller (1948), Einstein (1942, 1950) and Yalin (1963)
bed load equations all used some subset of the Gilbert (1914) and Meyer-
Peter et al. (1934) data to determine the coefficients in their equation, in-
cluding the D = 0.08 cm and D = 2.86 cm data of the respective studies.
(Einstein used only these two sets of measurements for his 1950 equation.)
117

J. Hydraul. Eng., 1989, 115(1): 101-123


Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

ioV

10c

10"1 ?_

4> /irnx
/r
11 jit.
10- 2 r //j* 10-2 u
/ / / • D= 0.079cm
/ / ^ * D=2.86cm
10" 3 ' '•' / Model (D= 0 . 0 8 c m ) _
Model(D=2.86cm)
k — Einstein ( 1 9 5 0 )

J. Hydraul. Eng., 1989, 115(1): 101-123


Meyer-Peter 81 Muller
L 1 1 1 i__i 1 1 in-*
2
10-2 2 4 6 e 10 -1 2 4 6 8iQ0 10" 2

FIG. 3. Comparison of Einstein and Meyer-Peter and Muller Bed Load Equations with Calculated Curves and Combined Data for D = 0.08
cm and 2.SS cm: (a) ib = -pghds/dx; (h) T„ = -pgR'ds/dx,
These two data sets can be fit together to produce a fairly smooth curve
covering a wide range of transport values as shown in Fig. 8; in doing this,
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

however, the grain size dependence is lost. The larger gravel data are all
for lower plane-bed transport while the smaller size includes transport over
bed forms and in upper plane-bed flow. Fig. 8(a) shows the data with T*
calculated in terms of h with Williams' side-wall correction; the values in-
dicated by the small symbols are Gilbert's bed form data, which could not
be corrected using Eq. 7. Fig. 8(b) shows the data with T* calculated in
terms of Einstein's (1950) sidewall- and bed-roughness-corrected R; here the
bed-roughness correction was applied to all values, including plane-bed mea-
surements, as Einstein appears to have done; the results when the Meyer-
Peter and Miiller corrections are used are almost identical.
The curves predicted by the model for both grain sizes shown in Fig. 8(a)
indicate that the transport relationships for these sizes are considerably dif-
ferent. Curves calculated using Yalin's bed load equation for two sediment
sizes agree very well with the model results. As expected, the Einstein curve
lies closer to the data, particularly the 2.86 cm gravel values, in Fig. 8(£>)
where the stress is calculated from the corrected hydraulic radius; the same
is true for the Meyer-Peter and Miiller relationship. This good level of agree-
ment requires that the entire Einstein or Meyer-Peter and Miiller correction
(for side-wall effects and for bed-roughness, even for planar bed data) be
applied to the data. While this can be relatively straightforward for flume
data, the complications of natural systems, e.g., channel curvature and mul-
tiple bed form length scales, makes application of these shear stress correc-
tions problematic for field data. If we consider the Einstein and the Meyer-
Peter and Miiller curves in Fig. 8(a) in the range 4> = 10~4 to 10~3 where
the data for the two sizes approach the respective critical shear stresses, both
curves fall near the middle of the range of nondimensional critical shear
stress values [(T#)cr = 0.03-0.06] for grain sizes of fine sand and larger
(essentially any size that moves as bed load). Thus, these equations basically
represent average transport relationships.
There are two significant respects in which the bed load equations dis-
cussed above and our model do not reflect natural bed load transport con-
ditions. First, all of the bed load transport expressions included in this com-
parison relate bed load sediment transport to skin friction, i.e., the shear
stress at the actual bed. This is straightforward when the bed is flat, but
when ripples or dunes are present on the bed accurate estimates of bed load
flux can only be obtained after a correction is applied to the shear stress on
the mean bed (T = — pQhds/dx) in order to determine a spatially-averaged
estimate of the skin friction. It might seem more appropriate to take the bed
forms into account in the calculations and to relate transport rates directly
to the measured (uncorrected) shear stress. The difficulty with relating bed
load transport to this stress is that a unique relationship must exist betwen
the height and the wavelength of bed forms as a function of grain size and
shear stress, and this relationship must be known accurately. An attempt was
made during this analysis to construct such a relationship using the WES
data for D = 0.02 cm, 0.035 cm, and 0.05 cm, but even for this limited
range of sizes and flume conditions the dependence of ripple height and
wavelength on transport stage could not be generalized. Thus, until we have
a better understanding of the relationship between flow parameters and bed

119

J. Hydraul. Eng., 1989, 115(1): 101-123


form geometry, the best approach is to use skin friction as the appropriate
shear stress on which to base T* for bed load calculations,, and to use Eq.
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

7 in conjunction with measured (or estimated) bed form properties to go from


the total boundary shear stress to TS/.
The second way in which these bed load transport expressions differ from
those desirable for many natural problems is embedded in the assumption,
explicit or not, of uniform sediment size. One of our purposes in developing
the bed load transport model for uniformly-sized sediment is to construct a
foundation for a bed load transport model for sediment characterized by a
mixture of sizes. It would be impossible to address this more complicated,
but obviously important, problem without first understanding in detail the
mechanics of uniform sediment transport. This limitation is not unique to
our formulation; all of the other bed load transport equations discussed herein,
with the exception of Meyer-Peter and Miiller, employed only bed load
transport measurements of relatively well-sorted sediment (compared to many
natural grain-size distributions) to set their coefficients. The fact that data
for different grain sizes were combined in these investigations is not rele-
vant, since the sediment used in the measurements for each size was fairly
uniform.
Einstein (1950) did include an empirical correction to his uniform-size
sediment transport relationship to account for the effect a poorly sorted bed
has on the probability that sediment in the various size fractions would be
transported. Meyer-Peter and Miiller (1948) set their two coefficients with
data for both uniform and mixed-grain-size sediment. Neither formulation,
however, includes two important parameters: the critical shear stress for ini-
tial motion of each size fraction relative to the other sizes present on a poorly-
sorted bed, or a parameter, such as the ratio of settling velocity to shear
velocity, to determine the mode of transport, i.e., rolling, saltation or sus-
pension, for each size class in motion. A bed load transport model such as
ours directly calculates rolling and saltation, and the addition of vertical ve-
locity fluctuations to the model makes it possible to determine the conditions
for incipient suspension and the boundary conditions for suspended load cal-
culations. We recently have derived an equation for the critical shear stress
of sediment on poorly-sorted beds (Wiberg and Smith 1987) in preparation
for extending our model to the problem of mixed-grain-size sediment trans-
port.

CONCLUSIONS

The bed load transport model presented herein differs from existing meth-
ods of calculating transport in the bed load layer in that the formulation
proceeds directly from the equations of motion for a particle in a fluid and
it contains no coefficients set using bed load transport measurements. The
good agreement between the calculated values and data taken from Gilbert
(1914), Meyer-Peter et al. (1934), and "Studies of river bed materials" (1935)
suggests that the basic model accurately reflects the dominant features of the
mechanics of sediment moving as bed load, at least until the concentration
of sediment is sufficiently high that intergranular collisions become impor-
tant. Not only can sediment flux rates be predicted for a large range of
conditions, but the details of the saltation layer also can be examined with
the model. Accurate knowledge of saltation parameters is important for a
120

J. Hydraul. Eng., 1989, 115(1): 101-123


number of transport-related problems. For example, the roughness of a bed
over which sediment is moving is related to the height of the bed load layer,
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

and the lower boundary condition for suspended sediment calculations is a


function of the sediment concentration in the saltation layer. Both of these
quantities can be found using the model. While at present the model focuses
on transport of uniform sediment, it also provides the essential foundation
for a model of mixed-grain-size bed load transport.
The equations of Einstein (1950), Meyer-Peter and Miiller (1948), Yalin
(1963), and Fernandez Luque and van Beek (1976) relating bed load sedi-
ment flux to boundary shear stress all produce curves with a structure similar
to our calculated bed load-transport relationship, and generally lie in a rel-
atively narrow band close to the flat-bed and bed form corrected data. The
scatter of the measurements around the various curves indicates that the val-
ues calculated with our model and with Yalin's expression provide the best
argument with the observations when stress is computed in terms of depth;
this is primarily a consequence of the grain-size dependence of the critical
shear stress terms in these formulations. The accuracy of bed load flux cal-
culations depends critically on the accuracy with which the boundary shear
stress is known. If one's knowledge of the boundary shear stress is only
approximate, a simpler equation of the Meyer-Peter and Miiller type will
provide adequate estimates of transport rate. However, all of these bed load
transport expressions relate sediment flux to skin friction so that appropriate
values of shear stress must be employed when bed forms are present. Ac-
curate estimates of bed load transport over rippled beds can be made after
the shear stress is corrected to remove the effects of bed forms, using a
technique such as that described by Smith and McLean (1977).

ACKNOWLEDGMENTS

This work was supported by the National Science Foundation under Grant
OCE-8310712.

APPENDIX I. REFERENCES

Abbott, J. E., and Francis, J. R. D. (1977). "Saltating and suspended trajectories


of solid grains in a water stream." Philos. Trans. Royal Society London, Ser. A,
284, 225-254.
Brownlie, W. R. (1981). "Compilation of alluvial channel data: Laboratory and field."
Report No. KHr-R-43B, W. M. Keck Lab., Calif. Inst. Tech., Pasadena, Calif.
Einstein, H. A. (1942). "Formulas for the transportation of bed-load." Trans., ASCE,
107, 561-597.
Einstein, H. A. (1950). "The bed load function for sediment transportation in open
channel flows." Tech. Bulletin No. 1026, U.S.D.A., Soil Conserv. Service, 1-
71.
Fernandez Luque, R., and van Beek, R. (1976). "Erosion and transport of bed-load
sediment." J. Hydr. Res., 14(2), 127-144.
Francis, J. R. D. (1973). "Experiments on motion of solitary grains along the bed
of water streams."- Proc. Royal Society of Lond., Ser. A, 332, 443-471.
Gilbert, G. K. (1914). "The transportation of debris by running water." Prof. Paper
86, U.S. Geological Survey, Reston, Va.
Graf, W. H. (1971). Hydraulics of sediment transport. McGraw-Hill Book Co., New
York, N.Y., 139-151.
Guy, H. P., Simons, D. B., and Richardson, E. V. (1966). "Summary of alluvial

121

J. Hydraul. Eng., 1989, 115(1): 101-123


channel data from flume experiments, 1956-1961." Prof. Paper 462-J, U.S. Geo-
logical Survey, Reston, Va.
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

Meyer-Peter, E., Favre, H., and Einstein, A. (1934). "Neuere Versuchsresultate uber
den Geschiebetrieb." Schweizerische Bauzeitung, 103(13).
Meyer-Peter, E., and Miiller, R. (1948). "Formulas for bed load transport." Proc,
2nd Meeting Int. Association Hydr. Res., Stockholm, 39-64.
Owen, P. R. (1964). "Saltation of uniform grains in air." J. Fluid Mech., 20(2),
225-242.
Smith, J. D., and McLean, S. R. (1977). "Spatially averaged flow over a wavy
surface." / . Geophysical Res., 82, 1735-1746.
"Studies of river bed materials and their movement, with special reference to the
lower Mississippi River." Paper 17, U.S. Waterways Experimental Station, Vicks-
burg, Miss.
Yalin, M. S. (1963). "An expression for bed load transportation." J. Hydr. Div.,
ASCE, 89(HY3), 221-250.
Wiberg, P. L., and Smith, J. D. (1985). "A theoretical model for saltating grain in
water." J. Geophysical Res., 90, 7341-7354.
Wiberg, P. L., and Smith, J. D. (1987). "Calculations of the critical shear stress for
motion of uniform and heterogeneous sediments." Water Resour. Res., 23(8), 1471-
1480.
Williams, G. P. (1970). "Flume width and water depth effects in sediment-transport
experiments." Prof. Paper 562-H, U.S. Geological Survey, Reston, Va.
Wilson, K. C. (1966). "Bed load transport at high shear stress." J. Hydr. Div.,
ASCE, 92(HY6), 49-59.

APPENDIX II. NOTATION

The following symbols are used in this paper:

A = particle cross-sectional area;


b = channel width;
cD = drag coefficient;
Cs = volume concentration of sediment;
=
c* vertical weighing function for sediment concentration in the bed
load layer;
D = grain diameter;
FD = drag force;
FL = lift force;
Ff = gravitational force, (ps - p)gV;
Fv = relative acceleration (virtual mass) force;
9 = gravitational acceleration;
H = ripple height;
h = flow depth;
qs = volume flux of bed load sediment;
R = hydraulic radius;
S = excess boundary shear stress, (T6 — Tcr)/Tcr;
s = water surface elevation;
T* = transport stage, T 4 /T„.;
t = time;
u = velocity vector;
u = component of velocity in the downstream direction;
«A = relative velocity between the fluid and sediment, Uf — Uj;
H* = shear velocity ( = V T / P ) ;
122

J. Hydraul. Eng., 1989, 115(1): 101-123


V = particle volume;
z = height above the bed;
Downloaded from ascelibrary.org by UFRJ - Universidade Federal Do Rio De Janeiro on 05/30/23. Copyright ASCE. For personal use only; all rights reserved.

Zo = bottom roughness parameter;


z* = matching height for form-drag correction;
a = geometric parameter, AD/V;
5fi = trajectory height;
•n = mean level of the bed surface;
K = von Karman's constant, =0.407;
\ = ripple wave-length;
V = kinematic viscosity;
P = fluid density;
P» = sediment density;
Ti = boundary shear stress;
Tcr = critical boundary shear stress for initial motion;
T„ = shear stress of the outer flow;
V = skin friction;
T* = nondimensional shear stress, T/[(P^ — p)gD]; and
nondimensional sediment transport rate, qs/[(ps/p ~ l ) ^ 3 ]
-©-

Subscripts
avg = average value;
/ = fluid;
s = sediment;
D = drag; and
L = lift.

123

J. Hydraul. Eng., 1989, 115(1): 101-123

You might also like