Sie sind auf Seite 1von 38

LECTURE SECTION 3

Sediment Transport
Sediment Transport Principles
Initiation of Motion of a Particle
Gravitational Force
Fluid Forces
Initiation of Bed Movement
Moments
Shields Parameter
Reference and Visual Assessment of First Motion
Deposition / Settling Velocity of Particles in Still Water
Stokes Law
Inertial Law
Suspended Sediment Transport
Significance of Suspended Load
Suspended Sediment Diffusion
Concentration Profiles and the Rouse Number
Sediment Transport Equations
Types of Bed Load Transport Equations
Excess Shear Stress [qb = f(b - c)]
Stochastic (i.e. probabilistic)
Stream Power [qb = f(Q, S) or f( Q/W, S)]
Empirical Correlation
Summary and Comparison of Bed Load Transport Equations
Approaching a Sediment Routing Calculation
Flow in Bends
The radius of curvature
Superelevation
Consequences for sediment transport

3-1

Sediment Transport Principles


The ability of the channel to entrain and transport sediment depends on the balance between (1)
gravitational forces acting to settle particles on the bed and (2) drag forces that act to either
suspend them in the flow or shove them along downstream.
Consider the forces acting on a grain on the bed surface ...
Assuming that the sediment is cohesionless, opposing forces act on the grain: the submerged
weight of the grain acts to hold it on the bed; and fluid lift and drag forces act to lift, roll, and
slide the grain along the bed. All of these forces are highly variable in nature and the general
problem of sediment transport is an incredibly messy affair -- complex enough to scare off the
greatest mind of this century ...

"As a young man, my fondest dream was to become a geographer. However


while working in the customs office I thought deeply about the matter and
concluded it was far too difficult a subject. With some reluctance, I then turned
to Physics as a substitute."
- Albert Einstein (Unpublished Letters)

Needless to say a few simplifying assumptions are in order:

We will consider only the simplest case of steady uniform flow over a flat bed of well sorted
grains.

3-2

Shields (1936) considered the problem of initiation of motion of a particle with diameter D lying
on a bed of identical grains.
There are 3 ways that grains can move:
1

lifting of the grain off the grains beneath it

sliding of the grain up and out of its position on the bed

rotation of the grain about a pivot point formed by neighboring grains

For case 1, lift forces must exceed the gravity force (i.e. the weight, since gravity itself is an
acceleration)
For case 2, drag forces in the direction of easiest movement must exceed the combined frictional
and gravitation force in the opposite direction
For case 3, the moment of the fluid forces must exceed the moment of the gravitational force.

3-3

Note that the gravitational force acts through the center of gravity, which generally will be close
to the center of the grain.
In contrast, the fluid force does not act through the center of the grain, because the position of
the pivot (and the direction of easiest transport) will vary from grain to grain depending upon the
relation to neighboring grains.

Forces exerted by the fluid vary from grain to grain due to relative exposure to the flow and due
to turbulent velocity fluctuations.

Due to such variation, any criterion that we can establish for the onset of sediment transport will
not be completely deterministic, but will be stochastic (i.e., statistical). Consequently, in
evaluating and applying models for the onset of sediment transport, we need a clear definition of
what is meant by the beginning of sediment movement (a.k.a. incipient motion).

3-4

Gravitational Force
The submerged weight of a sphere (Fg) is given by
Fg = mass * gravity = volume * density * g

(1a)

Fg = (4/3) (D/ 2)3 (s ) g

(1b)

By rearranging terms this can be expressed as


Fg = (/6) (D)3 (s ) g

(2)

The part of this force that opposes sliding of the grain in direction of easiest movement is equal
to
Fg sin

(3)

where is the friction angle of the material on the bed, which takes into account both the
geometry and the frictional properties of the bed material.
is generally about 35 for sand submerged in water.
For large grains on the bed, is smaller than for the average grain size, and for small grains is
larger than the average:

Large grains may actually be the easiest to move!


Due to this effect, incipient motion is often calculated for the median bed surface grain size (d50
or D50, depending on notation) to characterize general bed mobility.
3-5

Fluid Forces
A flowing fluid exerts 2 kinds of forces on the bed: drag and lift
1.

viscous drag on the upper exposed surface of the grain acts parallel to the bed

"Drag" is the friction due to the flow moving past the exposed grain surface. It therefore is a
function of: 1) surface roughness of the grain; 2) flow velocity; and 3) the exposed cross-sectional
area of the grain that the flow hits.

The standard drag equation has the form:


FD = CD ( u2/ 2) A
Where:

and

(4)

FD is the drag force


CD is the drag coefficient
u is flow velocity
A is the cross sectional area of the grain exposed to flow.

In general, we have the problem of what to choose for the drag coefficient CD.
It is also not clear what is appropriate for the velocity and exposed area -- think about a whole
bunch of weird-shaped grains sticking up into flow, whose velocity varies with height above the
bedan analytical nightmare.
Nonetheless, Coleman (1967) showed that the general form of (4) is true for spherical grains
sitting on a bed of identical grains.
3-6

2.
Lift forces due to unequal distribution of dynamic pressure on the grain surface act
normal to the bed
Higher fluid velocities develop over the top of the grain than underneath it, which gives rise to
higher pressures on the lower surface of the grain than on the upper surface, which results in a
net upward force:

The lift force is described by an equation similar to that for the drag force:
FL = CL ( u2/ 2) A
Where:

and

(5)

FL is the lift force


CL is the lift coefficient
u is flow velocity
A is the cross sectional area of the grain exposed to flow

Again, CL is only derivable for very simple cases and therefore must be experimentally
determined for natural stream beds.
Experiments have shown that static lift forces can be as much as 80% of the drag force at the bed
surface, but that lift forces decrease rapidly with height and are negligible at a few grain heights
above the bed.

3-7

Initiation of Bed Movement


An individual grain begins to move when
FD + FL > Fg

(6)

It is impossible to define this balance for all grains on a stream bed:

some grains lie in positions from which they are more easily lifted, slid or rolled

some grains are more exposed to the flow and subjected to larger fluid forces

fluid forces fluctuate with time because of turbulence in the flow

Again, we have to simplify:


Considering only drag forces and that friction prevents sliding of grains up and over each
other simplifies the problem to case # 3 above -- that for pivoting a grain up and out of its
position on the bed.
Total moment = product of the total force and distance to the pivot point

Driving :
Resisting :

a1 (FG sin )
a2 (FD cos )

The condition for the onset of motion is given by


a1 (FG sin ) = a2 (FD cos )

3-8

(7)

The gravity force may be written as


FG = c1 D3 g (s )

(8)

where c1 is a coefficient that accounts for non-uniform grain shape.


The fluid drag force may be assumed to be equal to the product of the average boundary shear
stress (b) and the area of the grain, which may be written as
FD = c2 D2 b

(9)

c2 is a coefficient that accounts for the geometry and packing of the grains (through the exposed
area) and for variation in the drag coefficient.
For initial motion, b = the critical shear stress, c, and so with substitution of equations (8)
and (9) into (7) we get:
a1 c1 D3 g (s ) sin = a2 c2 D2 c cos

(10)

Rearranging (10) in terms of c yields an expression for the critical shear stress required to
initiate motion:
c = (a1 c1 / a2 c2) tan D g (s )

(11)

Collecting all of the constants together into a single constant called *, the dimensionless critical
shear stress, or Shields parameter, this simplifies to
c = * D g (s )

(12)

Rearranging in terms of the constant yields


* = c / [D g (s )]

(13)

Note that * is the critical force divided by the submerged weight of the grain.

3-9

Because of all the uncertainties surrounding particle shape and the exact nature of grain-fluid
interaction, Shields (1936) conducted a set of laboratory experiments to investigate how *
would vary with the roughness Reynolds number (Re * = u*D/ ).

0.06

The left hand, low-roughness-Reynolds-number part of the curve shows that fine grains, hidden
partly within the laminar sublayer, would require a larger shear stress relative to their size than
would somewhat larger, more exposed particles. This range is important in the quieter parts of
large lowland rivers.
For most rivers, we are concerned with the right-hand side of the graph, where Shields data
roughly defines a dimensionless critical shear stress of 0.06.
Hence the initiation of bed movement by turbulent open channel flow is traditionally evaluated
using
c = 0.06 D g (s )

(14)

This relation was widely regarded as adequate for sand-bed rivers and little more was done until
the late 1970s when application of (14) to gravel-bed rivers was thought to be inaccurate because
the wider range of particle sizes on such beds causes a wider range of friction angles. Other
complications due to particle shape, degrees of protrusion, and orientation were also thought to
make Shields' value of 0.06 for * inappropriate for gravel-bed rivers.

3-10

Analyses of the case of non-uniform grain sizes have investigated the effect of changes in the
friction angle and grain protrusion into the flow. The "cleanest" such studies tend to yield
Shields-like results with the important difference that * is a strong function of grain size relative
to the grains that compose the stream bed.
Big particles can move over small particles at relatively low c
Small particles require a relatively higher c to move them across beds of larger particles.

And, as an added complication, the % of sand in a mixed-sediment bed can profoundly alter the
rate of sediment transport (see Wilcock et al., 2001, Experimental study of the transport of
mixed sand and gravelWater Resources Research, 37(12):3349-3358increasing sand content
from 6 to 21% increased sediment transport by 1 to 3 orders of magnitude for the same water
discharge!). Thus, it has proven surprisingly difficult to quantify the critical shear stress for
motion of mixed-grain-size, coarse streambeds.
Two methods of defining the initiation of motion have been commonly used:
1.
Direct observations of the first movement as shear stress increases. This is notoriously
difficult to do, even in a laboratory flume.
2.
Measurement of bed load transport rates (qb) and plotting them against basal shear stress
(b). Extrapolation of the resulting trend to a value of qb = 0 has been the preferred method,
either in the field or lab, but you can only use this for the average particle size.

qb

3-11

Empirically derived * values have distinct biases due to differences in experimental or


observational methodology (Buffington and Montgomery, Water Resources Research, 1997):
Visual-based (i.e. visually observed determination of first transport).

Oak Creek

Reference-based (i.e. project qb to condition of zero transport).

So what value of * do you use? In part, the range of plausible values should correspond to the
range expressed in the graphs above for your particular applicationis the problem one of first
motion? Or is it that of net bed load transport? In the summary of Buffington and
Montgomery (1997):
Our reanalysis of incipient motion data for bed surface material indicates that (1) much of the scatter in
Shields curves is due to systematic biases that investigators should be aware of when choosing and
comparing dimensionless critical shear stress values from the literature; and (2) there is no definitive *50
value for the rough, turbulent flow characteristic of gravel-bedded rivers, but rather there is a range of
values that differs between investigative methodologies. Our analysis indicates that less emphasis should
be placed on choosing a universal *50 value, while more emphasis should be placed on choosing defendable
values for particular applications, given the observed methodological biases, uses of each approach, and
systematic influences of sources of uncertainty associated with different methods and investigative
conditions.

3-12

Deposition / Settling Velocity of Particles in Still Water


Transport of grains occurs when instantaneous vertical component of flow exceeds the settling
velocity of particles.

Fall velocity in still water is influenced by two sets of forces:


1
submerged weight of particle
2
viscous fluid resistance and inertia effects
Small and big particles behave differently:
for small particles, viscous resistance dominates; inertia is negligible.
for large particles that fall quickly, inertial forces dominate.
Conditions under which of these effects dominates can be discerned using a particle Reynolds
number (Rep)
Rep = 0 D /

(15)

where 0 is the fall velocity, D is the particle diameter, and is the kinematic viscosity of the
fluid.
Based on experiments, inertia is negligible where Rep < 1: the boundary layer around the particle
is laminar, and the fall is smooth (silt and clay). For Rep > 100, the viscous force is negligible,
the boundary layer around the particle is turbulent, fall is rough, and a wake develops (gravel).

3-13

Stokes Law
Stokes (1851) considered the problem of the balance between the downward force due to the
submerged weight of a particle and a viscous resistance force and thereby got a "law" named after
himself.
Submerged weight of a spherical particle
W = (/6) D3 (s ) g

(16)

Hopefully, this looks familiar [recall equation (2)]

Viscous resistance = f ( surface area, dynamic viscosity, and fall velocity )


V = 3 D 0

(17)

where = dynamic viscosity and 0 = fall velocity.

Assuming no acceleration of a falling particle (i.e., constant terminal velocity, s) then W = V


and
(/6) D3 (s ) g = 3 D s

(18)

Rearranging (18) in terms of the fall velocity yields


s = (1/18) D2 (s ) g /

(19)

Hence the fall velocity of small particles (i.e. with no inertial effects; see eqn. 17) is proportional
to the square of the diameter.
Note also that the settling velocity depends upon the viscosity of the fluidcolder temperatures
or high sediment concentrations can make larger particles behave like finer ones and settle
slower, thereby allowing smaller upward velocity fluctuations to keep them suspended above the
stream bed.
3-14

Inertial Law
Particles larger than about 2 mm (i.e., sand) encounter resistance from the "impact" force given
by the momentum per unit time of the cylindrical column of water whose cross-section area is
the projected area of the falling grain.
The impact force is given by
I = (1/4) D2 s2

(20)

The force balance between the submerged weight of this falling particle and the impact force is
given by equating (16) and (20), at which time o = s, the terminal settling velocity:
(/6) D3 (s ) g = (1/4) D2 s2

(21)

Rearranging terms yields:

s =

( )
2
Dg s
3

(22)

Hence, large grains fall at a rate proportional to the square root of their size, which leads to a
strong size dependence to the fall velocity of particles in streamflow:

Whether particles of a particular size stay suspended in the flow or settle to the bed depends on
the magnitude of upward turbulence.
3-15

Suspended Sediment Transport


Although the morphology of many river beds is set by bed load transport, the majority of
sediment moved by rivers travels as suspended load.

Significance of suspended load

accounts for 80 - 95+% of sediment flux to oceans

flood plain formation: flux of sediment to flood plain = suspended load + overbank
discharge

At low slopes and discharges, a suspended sediment layer can develop on stream beds and
be incorporated into streambed gravels

Many water-quality contaminants travel as adsorbed constituents on the surface of


suspended sediment particlesso water-quality problems commonly can be analyzed only
by understanding the suspended-sediment transport

The total suspended sediment load is normally described as:


z=H

Qsusp. =

C(z) u(z) dz

(23)

z =0

where C(z) is the concentration profile of the suspended sediment.

What is the distribution of suspended sediment within the water column?

Particles settle via gravity to the bed at a terminal settling velocity, s. As particles settle, a
concentration gradient develops, with more sediment deeper in the flow. However, the upward
component of turbulence results in a mass flux up into the flow column that acts to maintain
suspended sediment transport. The balance between downward settling and upward diffusion is
what establishes the equilibrium concentration gradient and provides a means to develop a
mathematical expression for that concentration.

3-16

Suspended Sediment Diffusion


Two Processes
1

Settling by gravity occurs, with a terminal settling velocity of s.


The downward flux of mass per unit area of a plane parallel to the bed is
Fdown = Cs

(24)

where C is the volumetric concentration of sediment (e.g., in units of kg/m3). This


process concentrates sediment towards the bed of the channel.

Eddy exchange between layers in the turbulent flow causes a net flux of sediment
between layers with a difference in sediment concentrationrandom turbulent
interactions between layers cause intermixing at all levels, and therefore a net transport
from areas of high concentration into areas of lower concentration. This net flux per unit
area will tend to lift suspended sediment and is described by:
Fup = Ks (dC/dz)

(25)

where Ks is the sediment mass diffusivity. The negative sign results from flux down the
concentration gradient (i.e. upward in the flow).

3-17

The sediment mass exchange by turbulence is a diffusion process (the flux is proportional to a
potential gradient, in this case the concentration of suspended sediment). This process is
analogous to the exchange of momentum that generates an eddy viscosity in the derivation of
velocity profiles in turbulent flow (see the notes on the derivation of the Law of the Wall). In
the case of momentum exchange, the flux of momentum per unit area of horizontal plane was
the means by which a shear stress was exerted between layers. We first wrote that equation as
= ( + )(du/dz) (Equation 2-31), for which we then argued it could be adequately
represented by = (du/dz). Here, we will define the eddy diffusivity as Km = /, and so:
= Km (du/dz)

(26)

where Km is the diffusivity of eddy momentum in the flow.

Vanoni made the analogy between the sediment mass diffusivity, Ks, and this corresponding
water eddy momentum diffusivity, Km. He said that
Ks Km

(27)

where should be less than 1 for coarser particles and converges on 1.0 for fine particles.
In other words, = 1 if the movement of sediment just tracks the movement of water, but larger
particles can't keep up, so < 1 for them.

(Please forgive the lack of any equations numbered 28 or 29)

3-18

Recalling the derivation of the Law of the Wall, Prandtl argued that near the bed, b (the
shear stress on the bed) and that Km = k u* z. In this equation, k is Von Karmans constant
(0.4).
With these assumptions, equation (26) can be re-written as:
/ = Km (du/dz) = k u* z (du/dz)

(30)

Recall also from fluid mechanics that


u* =

or

b / =u*2

(31)

Substitution of equation (31) into equation (30) results in


(du/dz) = u* / kz

(32)

Returning to equation (26) and by analogy to the Law of the Wall, at any level in the flow:
/ = (b /) [(H z)/H]

(33)

where H is the flow depth.

Combining (33) and (31) results in:


/ = u*2 [(H z)/H]

(34)

Substitution of equation (26) into the left hand side of equation (34) yields:
Km(du/dz) = u*2 [(H z)/H]

3-19

(35)

Similarly, substitution of equation (32) into equation (35) yields


Km u*/kz = u*2 [(H z)/H]

(36)

which can be rearranged to yield


Km = u* kz [(H z)/H]

(37)

This expression for the water eddy momentum diffusivity (Km), together with the assumption
that Km Ks for fine particles (recall 1.0), allows us to return to equation (25), the expression
for the upward flux of suspended sediment due to turbulent diffusion:

Fup = Ks(dC/dz)

(38)

= Km(dC/dz)
= k z u* [(H z)/H] (dC/dz)

Now we can consider the situation in which an equilibrium has been established between the two
vertical sediment fluxes (i) settling of particles and (ii) turbulent exchange of sediment mass.
At this point (i) = (ii), Fup = Fdown, and
C s = k z u* [(H z)/H] (dC/dz)
where C is varying with z; i.e., C = C(z).

3-20

(39)

To solve, we first separate variables in (39):


dC / C = [s / k z u*] [H /(H z)] dz

(40)

The solution of equation (40) is one of those for which the labor of finding does not enhance the
utility of the result. So take it on faith (but check if you must) that:
ln C = [s / k u*] ln[H /(H z)] + C

(41)

where C is the constant of integration. If we choose the value of C cleverly, this equation can
be rewritten as:
ln C = ln Ca + [s / k u*] ln{[H /(H z)][a/(H a)]}

(42)

where Ca is the concentration at some (any) one elevation above the bed at which z = a.
Finally, this can be simplified as:
a
H z

C ( z) = Ca
H a
z

u* k

(43)

assuming that = 1. This more clearly expresses the form of the predicted relation for the
concentration profile of suspended sediment as a power function of the distance from the bed
surface.
Note that in equation (43), the term a/(Ha) is less than 1 in the lower of the flow (where,
typically, any significant concentration of suspended sediment will be measured). The product
with the term (Hz)/z is also less than one over most of the flow depth, and so raising it to an
exponential power much greater than one will drive the value of C(z) towards zero. Thus the
maximum value of the suspended sediment concentration will be determined by the empirically
fit magnitude of Ca (and where in the flow that value was measured), but the rapidity with which
that concentration gradient declines with distance away from the bed will be determined by the
magnitude of the exponent, s/ku* the larger this number (i.e. the coarser the sediment, or the
lower the shear) the more abruptly will the concentration decline away from the bed.

3-21

This variable, s/ku*, is dimensionless (velocity velocity) and is known as the Rouse number
(p = s/ku*).

In general, values of the Rouse number greater than about 2.5 lead to a condition of very little to
no suspended sediment. In contrast, values of the Rouse number less than about 0.25 predict
those grain sizes that will move as wash load, fully supported by the flow.
In between these values lies the range of partial suspended sediment transport. A Rouse number
less than 1.8 is sometimes suggested at an appropriate functional threshold for determining
whether a given grain size will move in suspension (but note that the process itself is not
threshold-driven!).
Note also that smaller particles are predicted to be relatively evenly distributed through the flow
column and that larger particles should be found only close to the bed. Although the form of
equation (43) requires that the suspended sediment concentration be zero at the top of the water
column (where z = H) regardless of particle size, we know from observations that this is not
always the case. Which assumption(s) of the mathematical development is (are) not met in
practice?

3-22

Ca in equation (43) can be estimated in three ways:

1. By fitting the equation to sediment concentration profiles measured with point samplers at
various elevations above the bed.

2. By back-calculation from the vertically averaged sediment concentration measured in a


depth-integrated suspended load sampler

3. By estimating the bed load flux (via equations or samplers), estimating the top of this bed
load layer as a in (43), and converting the bed load flux through depth increment 0 <z <a
into a concentration, Ca.

Where sediment movement in the flow column is determined by the Rouse number
(p = s/ku*), the following conditions are anticipated:
p > 2.5

Grains travel as bed load

1.8 < p < 2.5

Grains are trying to go into suspension, but are not quite able to stay up in
the flow field

p < 1.8

Full suspension

Hence, you can use the previous derivations of the settling velocity and shear velocity to
determine what grain sizes will travel as suspended or bed load based on the ratio of the settling
velocity (s) to the shear velocity (u*).

3-23

Recall from previously,


2

s = (1/18) D (s ) g /

(19)

u* = (b/)0.5

(44)

and by definition,

The largest size of suspended material can be determined from the condition of s = u*, which
implies a Rouse number equal to 2.5, i.e.
s/ku* = 2.5

Thus equating (19) and (44) and rearranging terms yields:


0.5

D2 = [18 (gHS) ]/(s ) g

A similar approach using p < 1.8 would solve for the largest size of wash load.

3-24

(45)

Sediment Transport Equations


We are going to focus on bed load transport (even though suspended load dominates total
sediment flux) because the beds of rivers are made of bed load sediment and the morphology of
rivers is set by bed load transport.

Bed load moves immediately above the bed in a rolling or saltating (jumping) mode.

Bed load is generally > 0.2 mm in diameter, although the size of material that travels in
suspension (suspended load) varies with the power and turbulence of the flow.

Bed load transport equations ignore the fact that grains move via turbulent bursts that must
penetrate the laminar sub-layer, which shields grains from turbulence and movement of the flow.
Because of this we back off from the force balance approach and develop relations between
observed sediment transport rates and mean measures such as Q, S, W, h (or d), D50, or b.

Each bed load transport formula has been developed and calibrated from specific conditions of b
S, D, and qbso the basic rule of thumb is to use the one whose conditions best approximate
your river.

Be smart about how you apply sediment transport equationsfor example, if field evidence
shows that transport only occurs in the middle of a river, then dont use mean flow depth based
on whole cross-section. Instead, use flow depth over the active transport zone or break the
channel into zones with different flow depths. Attempts to calculate sediment transport are
replete with cookbook application of formulas without any physical insight into the relevant
conditions. Do the geomorphology first, and then try to attach some numbers to it!

3-25

Types of bed load transport (qb) equations


Geologists and hydrologists have proposed numerous bed load transport equations; they fall into
four general types:

1.

Excess Shear Stress [qb = f(b c)]

2.

Stochastic (i.e. probabilistic)

3.

Stream Power [qb = f(Q, S) or f( Q/W, S)]

4.

Empirical Correlation

1. Excess Shear Stress


Most bed load transport equations represent qb as the difference between the applied and critical
parameters required for initiation of grain mobility:
qb (b c)

(46)

where b is the effective boundary shear stress (i.e. applied to the bed sediment), c is the critical
boundary shear stress required for grain movement, and is an empirical exponent, typically >
1.
a)

Du Boys (1879)

The oldest approach to excess shear stress:


qb = b (b o)

(47)

where = (0.173)/(D50)3/4 with D50 in mm, and o is almost (but not exactly) Shields c
(which, recall, is a function of D50).

3-26

b)

Meyer-Peter equation

An empirical relation based on a large number of experiments with uniform sediment 329 mm,
in flumes 20200 cm wide, and depths < 20 cm. So: coarse sand, no bedforms, and grain
roughness total roughness.
The equation is traditionally (and somewhat confusingly) expressed in different units, depending
on the system of measurement:
For the SI system, the units of qb (sediment discharge) are kilograms of sediment
transported per second per meter of channel width; q (water discharge) is in units of cubic meters
per second per meter of channel width; and D50 (median sediment diameter) is in units of meters
(not mm!).
For the English system, the units of qb are pounds of sediment transported per second per
foot of channel width (note that this is a unit of weight, not mass as for the SI version); q is in
units of cubic feet per second per foot of channel width; and D50 (median sediment diameter) is
in units of feet.
The equation is:
SI units:

English units:

qb = [250 q2/3 S 42.5 D50]3/2

qb = [39.25 q2/3 S 9.95 D50]3/2

(48a)

(48b)

Historically, the MP equation has been used to evaluate the bed load transport rate of different
fractions of the bed load sediment population by substituting different values for D50 and
multiplying the resulting qb by the fraction of total bed sediment represented by this size range.
This is an effective way to investigate the routing of different sediment sizes down a channel.
Unfortunately, such an approach has no theoretical justification and, in fact, has many reasons
for why it is not correct. Such details have not stopped this practice!

3-27

c)

Meyer-Peter & Muller (1948)

An empirical relation based on a large number of experiments with uniform, graded, and
lightweight materials; a more theoretically based expansion of the original Meyer-Peter equation.
They formally suggested that a single effective grain diameter (D50) be used to characterize mixed
sediment and that the total shear stress be modified to account for grain roughness, reducing the
shear stress available for transporting sediment.
Their empirical data were still flume-based but with a wider range of variables: grain sizes
between 0.4 and 30 mm, depths between 1 and 120 cm, and a ratio of grain roughness to total
roughness of 0.5 to 1.
Their equation is:
n
s
1

2
8 grain roughness
q b =

ntotal roughness
s

2
b 0.047( s )gD50

(49)

qb is in units of weight per second per unit width of channel, and the entire equation is
dimensionally homogeneous (so we dont need to worry about which measurement system we are
usingjust be consistent!). Note that the last term in the brackets looks very much like Shields
equation for c with a slightly lower value for * (0.047 instead of 0.06). Note also that when this
term exceeds the b term, transport isnt negativeit just doesnt happen because the threshold
of motion is presumed to have not yet been reached.
s and are densities of sediment and water (2650 and 1000 kg/m3), g is gravitational
acceleration (9.81 m/s2). Grain and total roughnesses are represented with Mannings n, where
the grain roughness is defined by the Strickler (1923) equation:
ngrain roughness = 0.0151 D501/6

(50)

where D50 is in mm, and ntotal roughness is determined by any of the methods normally used to
estimate Mannings n.

3-28

d)

Parker (1990)

Another empirical equation, based on analysis of data from Oak Creek, Oregon, a small paved
gravel-bed channel. You will note that the equation does not specify a threshold for incipient
sediment transport. Their analysis assumes subsurface sediment is the source of bed load, but
this can occur only once the bed surface pavement is breached.
Define first a dimensionless bed load transport rate, w*:

qb s

w* = 1
3
g 2 (dS ) 2

(51)

or,
qb = g

3
(dS ) 2

w*
s

(52)

Note that this is also dimensionally homogeneous, but that qb is in units of volume of sediment
per second per unit width .
Now the only problem is to determine w*. This is where the empirical data from Oak Creek
came in: Parker expressed the data in two ranges, depending on the value of the dimensionless
shear stress of the stream as expressed by yet another dimensionless parameter, 50:

50 =

where

b
gdS
0.0876 =
0.0876
( s )gDspvt 50
( s )gDspvt 50

(53)

Dspvt50 is the median diameter of the subpavement, not the surface sediment.

We can now solve Parkers equation for qb by determining w*:


w* = 0
w* = 0.0025 exp[14.2(50 1) 9.28(50 1)2]
w* = 13.69 [1 (0.853/50)]4.5

3-29

for 50 < 1
for 1 < 50 < 1.6

(54a)
(54b)

for 1.6 < 50

(54c)

Stochastic Approach
The Einstein (1950) bed load transport formula is the most complex procedure developed for
natural streams.

It is a probabilistic approach, based on sand bed rivers in the Mississippi basin. He assumed that
the number of particles eroded per unit time equals the number of particles on the bed times the
probability of any one being eroded in an a unit time. We can consider sediment transport in 2
ways: the number of grains leaving a unit area in a unit period of time, and the number of grains
entering an area in a unit period of time. The former (leaving grains) is expressed in terms of the
probability of entrainment and the settling time of a grain once entrained (notice this assumes
that saltation is the mode transport); the latter (entering grains) is expressed in terms of the bed
load transport rate (what we will eventually want to solve for) and the length that a saltating
grain will hop.
At steady state, these two rates are equal. The probability of entrainment is a function of the lift
forces and the buoyant weight of the particles. It has been determined by experiment and is
expressed as a flow-intensity index. With corrections for grain hiding and hydraulic roughness,
a transport-intensity function is read off a graph relating it to the corrected flow-intensity index
and then converted into a bed load transport rate for the given particle size of interest. As with
the Meyer-Peter equation, this equation is commonly run for discrete grain sizes under the
assumption that each moves independently of the others. Insofar as this is a saltating system,
presumably without a pavement layer, this assumption may not be as ill-conceived as in gravelbed systems.

3-30

Stream Power Approach


Bagnold (1960, 1966) developed an approach to bed load transport prediction that is based on
stream power per unit channel width (), which is defined as:
= g (Q/w) S

(55)

He argued, reasonably, that was an appropriate measure of transport. Recall, from classical
mechanics, that
Work = force x distance
Power = work/unit time = force x distance/time = force x velocity
And because stress = force/area,
Power/area (Bagnolds , the unit stream power) = stress x velocity
He also defined the power of the moving sediment per unit bed area as:
(s )g (Vseds / Vbed) tan useds
where Vseds / Vbed is the concentration (by volume, V) of sediments on the bed. The entire first
term of this expression looks very much like the frictional strength part of the slope-stability
equation of classical landslide mechanics with , Bagnolds slipping angle, playing the role of an
angle of internal friction.
Bagnolds key assumption is that stream power is transmitted to moving sediment via an
efficiency factor, eb, that ranges from 0 to 100%. With a value of 0 there is no sediment motion;
at 100%, there would be no dissipation of energy in the system at all except from the movement
of grains (which is obviously impossible). USGS Professional Paper 422H tabulates presumed eb
values as a function of mean water velocity.
The method as a whole has never seen much application except in the analysis of very sedimentrich flows (such as debris flows).

3-31

Empirical Correlations
Bagnold was convinced that stream power was the correct theoretical approach to the problem of
sediment transport, but assessing the efficiency factor was very problematic and the focus of
many subsequent papers. After nearly 2 decades, Bagnold gave up. Instead, he decided that
even if an analytic form could not be defined, there surely were enough data to find an empirical
relationship between unit stream power and sediment transport. In recognition of other
potentially influential factors, he also included flow depth and grain size of the bed load in his
final correlation.
His equation was expressed in terms of reference conditions (marked with an asterisk, and so
given values) and the subject conditions (unasterisked, user-supplied). The three terms of the
equation are for excess unit stream power, flow depth, and modal grain size (normally, median
can be substituted here). The equation is:
3

o 2 Y
ib = ib*

( o )* Y*

D

D*

(56)

ib is in units of (watch closely!) the submerged mass rate of sediment transport per unit width of
channel. To convert to dry mass, multiply by [s/(s-)], or about 1.6. To convert to weight,
multiply by g (i.e. 9.8 m/sec2 in SI units).
Reference (i.e. *) values are as follows:
ib* = 0.1 kg/m-sec
( - o)* = 0.5 kg/m-sec
Y* = 0.1 m
D* = 1.1 x 10-3 m (about 1 mm, but remember that the units are METERS!)
For your stream, Y = average flow depth and D = modal grain size (both in m).
is the unit stream power, calculated from the definition (equation (55) above). o is the unit
stream power at the threshold of motion; Bagnold assumed a Shields parameter of 0.04 and
substituted that into the definition of as the product of shear stress and mean velocity to get:
o = 290 D1.5 log[12Y/D]

3-32

(57)

Summary and Comparison of Bed Load Transport Equations

Each equation has a particular range of best applicability, a


function of both the assumptions and the size of the sediments and
flow conditions used to extract the empirical coefficients (a.k.a.
fudge factors). In particular, those with explicit data sets have
an obvious established range; Bagnolds original stream-power
approach required very concentrated flows to work well, and the
stochastic approach requires abundant saltating grains (i.e. sand)
and good measurements.

Accuracy is a tremendous problem with these equations.


Methods may vary by an order of magnitude (or more) in their
predictions. Sometimes as good an estimate can be made by
assessing the basin sediment yield and assuming that 1-10 percent
of that will travel as bed load. Also beware of supply-limited
systems, which violate assumptions of these methods.

See:
Gomez and Church, Wat. Res. Research, v. 25(6), p. 1161-1186
Most sediment transport formulae characteristically overpredict
sediment flux by 2-10 times due to:

failure to include surface coarsening

variations in the rate at which material is supplied and


available for transport
Best equations in general appear to be Bagnold (for gravel-bed rivers)
and Einstein (for sand-bed rivers).
Parker equation is a great favorite among theoretical geomorphologists
and is good most of the time, but at other times gives really wacky
answers

3-33

So, how to approach a sediment routing calculation?


1.

Sediment Supply

2.

Is bed material bed load or suspended load?


Is there another source of bed load that moves over an immobile substrate?
Need to determine source and texture of sediment for your problem

Scales of bed load movement

3.

Bankfull bed load movement


Low-flow fine sediment movement

Bed armoring

4.

pavement or no pavement

Which formula to use?


Consult Gomez and Church for equation that best fits your channel.

5.

Hydraulic parameters (u, H, S, & D)

How do you get these values and are they realistic?

Where do you sample D in a channel?

Choose the obviously mobile patches and the most representative.

Do calculations for simple, straight, plane-bed reach if possible as you can get a valid
average flow depth (H).

Dont always cross-section average - i.e., dont use average depth in a non-rectangular
channel.

3-34

Flow in Bends
The radius of curvature
Definition: the radius of the arc that traces out the meander path (not really circular, but thats
ok).
By observation, many rivers have a ratio rc/w = 2 or 3; studies of channel migration
rates suggest that rivers migrate more rapidly as this ratio increases (i.e. straight rivers migrate
fastest).
Why is this? Studies on pipe flow suggested that frictional resistance was lowest in this
range; other flume studies note that at radii much different from this range, either the upstream
limb migrates more rapidly than the downstream limb (too large) or visa versa (too small). Either
way, the result is convergence on this rangeso its not that the river somehow prefers this
curvature, its just that other curvatures are less stable on the landscape.

3-35

Superelevation
The centrifugal force of the water accelerating around the bend must be balanced by
something (or else it would continue to climb ever-higher)its the cross-stream pressure
gradient imparted by the superelevation (h):

u 2 h
rc

= gh

h
w

(58)

u2w
rc g

(59)

And so:
h =

Note the typical scale of the superelevation: rc / w is about 2 or 3; for a river with a flow of
about 1 m/s, h is a few cm.
The pressure gradient balances the average flow velocity, but near the surface the local
velocity exceeds the average velocity and so the centrifugal force exceeds the pressure; at depth,
average velocity exceeds the local velocity and so the pressure exceeds the centrifugal force:

So bottom flow is towards the inside of the bend. This is secondary flow (primary flow is
in the downstream direction).

3-36

3-37

Consequences for sediment transport


First, note the basic conditions that establish the spatial patterns of sediment erosion and
deposition:

Erosion occurs where b is increasing in the downstream direction.


Deposition occurs where b is decreasing in the downstream direction.
You wont find grains of a particular size unless they can be transported to their point of
deposition. So, in general, you will find the coarsest bed particles along the path of max (or
where max was when the channel was at a sediment-transporting stage!).

In addition, bedforms tend to channelize the secondary flow (because it moves along the
bed), and so the idealized pattern can be modified by the sedimentary deposition that results
from it: the flow shapes the bed, but the bed alters the flow. Any such system presents
opportunities for both negative feedback (i.e. stable channel form) and positive feedback (i.e.
rapid instability and migration).

3-38

Das könnte Ihnen auch gefallen