6.sedimentation and Erosion Hydraulics

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 102

6.

SEDIMENTATION AND EROSION HYDRAULICS


Marcelo H. García

Department of Civil and Environmental Engineering

University of Illinois at Urbana-Champaign

Urbana, IL

6.1. INTRODUCTION
Since the beginning of mankind, sedimentation processes have affected water supplies, irrigation, agricultural practices, flood
control, river migration, hydroelectric projects, navigation, fisheries, and aquatic habitat. In the last few years, sediment also has
been found to play an important role in the transport and fate of pollutants; thus, sedimentation control has become an
important issue in water quality management. Toxic chemicals can become attached to, or adsorbed by, sediment particles and
then be transported to and deposited in other areas. By studying the quantity, quality, and characteristics of sediment in rivers
and streams, scientists and engineers can determine the sources of the sediment and evaluate the impact of pollutants on the
aquatic environment. In the United States, sedimentation control is a multibillion-dollar issue. For example, approximately $500
million are spent every year to dredge waterways and harbors for navigation purposes. Most of the dredged sediment is the
result of substantial soil erosion in watersheds. Estimates by the U.S. Department of Agriculture indicate that annual offside
costs of sediment derived from copland erosion are on the order of $2 billion to $6 billion, with an additional $1 billion arising
from loss in compared productivity.

The sediment cycle starts with the process of erosion, where by particles or fragments are weathered from rock material.
Action by water, wind, glaciers, and plant and animal activities all contribute to the erosion of the earth’s surface. Fluvial
sediment is the term used to describe the case where water is the key agent for erosion. Natural, or geologic, erosion takes
place slowly, over centuries or millennia. Erosion that occurs as a result of human activity may take place much faster. It is
important to understand the role of each cause when studying sediment transport.

Any material that can be dislodged is ready to be transported. The transportation process is initiated on the land surface when
raindrops result in sheet erosion. Rills, gullies, streams, and rivers then act as conduits for the movement of sediment. The
greater the discharge, or rate of flow, the higher the capacity for sediment transport.

The final process in the cycle is deposition. When there is not enough energy to transport the sediment, it comes to rest. Sinks,
or depositional areas, can be visible as newly deposited material on a floodplain, on bars and islands in a channel, and on
deltas. Considerable deposition occurs that may not be apparent, as on lake and river beds. A knowledge of sediment dynamics
is an integral part of understanding the aquatic ecosystem.

This chapter presents fundamental aspects of the erosion, transport, and deposition of sediment in the environment. The
emphasis is on the hydraulics of bedload and suspended load transport in rivers, with the goal of establishing the background
needed for sedimentation engineering. Because of their relevance, the hydraulics of both reservoir sedimentation and turbidity
currents also is considered. Emphasis is placed on noncohesive sediment transport, where the material involved can be silt,
sand, or gravel. When possible, the behavior of both uniform-sized material and sediment mixtures is analyzed. Although such
topics as cohesive sediment transport, debris and mud flows, alluvial fans, river meandering, and sediment transport by wave
action are not discussed here, it is hoped that the material covered in this chapter will provide a firm foundation to tackle
problems in those.

For more information on sediment transport and sedimentation engineering, readers are referred to Allen (1985), Ashworth et

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
al. (1996), Bogardi (1974), Bouvard (1992), Carling and Dawson (1996), Chang (1988), Coussot (1997), Fredsøe and Deigaard
(1992), Garde and Ranga Raju (1985), Graf (1971), Jansen et al. (1979), Julien (1992), Mehta (1986), Mehta et al. (1989a,
1989b), Morris and Fan (1998), Nakato and Ettema (1996), National Research Council (1996), Nielsen (1992), National
Research council (1996), Parker and Ikeda (1989), Raudkivi (1990, 1993), Renard et al. (1997), Sieben (1997), Simons and
Senturk (1992), Sloff (1997), van Rijn (1997), Yalin (1972, 1992), Yang (1996), and Wan and Wang (1994).

6.2. HYDRAULICS FOR SEDIMENT TRANSPORT


6.2.1. Flow Velocity Distribution
Consider a steady, turbulent, uniform, open-channel flow having a mean depth H and a mean flow velocity U (Fig. 6.1). The
channel is extremely wide and its bottom has a mean slope S and a surface roughness that can be characterized by an effective
height ks (Brownlie, 1981b). When the bottom of the channel is covered with sediment having a mean size or diameterD, the
roughness height ks will be proportional to that diameter. Because of the weight of the water, the flow exerts on the bottom a
tangential force per unit bed area known as the bed shear stress τb, which can be expressed as:

(6.1)

where ρ is the water density and g is the gravitational acceleration. With the help of the boundary shear stress, it is possible to
define the shear velocity u* as

(6.2)

Figure 6.1 Definition diagram for open-channel flow over an erodible bed.

The shear velocity, and thus the boundary shear stress, provides a direct measure of the intensity of flow and its ability to
entrain and transport sediment particles. The size of the sediment particles on the bottom determines the surface roughness,
which in turn affects the flow velocity distribution and its sediment transport capacity. Since flow resistance and sediment
transport rates are interrelated, the ability to determine the role played by the bottom roughness is important.

Research has shown (Schlichting, 1979) that the flow velocity distribution is well represented by:

(6.3)

where ū is the time-averaged flow velocity at distance z above the bed and κ is known as Von Karman’s constant and is equal to
0.4. For obvious reasons, the above law is known as the logarithmic law of the wall. It strictly applies only in a thin layer near the
bed. It is empirically found to apply as a reasonable approximation throughout most of the flow in many rivers.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
If the bottom boundary is sufficiently smooth (a condition rarely satisfied in rivers), turbulence will be drastically suppressed in
an extremely thin layer near the bed. In this region, a linear velocity profile will hold:

(6.4)

where v is the kinematic viscosity of water. This law merges with the logarithmic law near z = δv, where

(6.5)

denotes the height of the viscous sublayer. In the logarithmic region, the constant of integration introduced above has been
evaluated from data to yield

(6.6)

Most boundaries in river flow are rough. Let ks denote an effective roughness height. If ks/δv > 1, then no viscous sublayer will
exist. The corresponding logarithmic velocity profile is given by

(6.7)

As noted above, this relation often holds as a first approximation throughout the flow in a river. It is by no means exact.

The conditions ks/δv » 1 for rough turbulent flow and ks/δv « 1 for smooth turbulent flow can be rewritten to indicate that u*ks/v
should be much larger than 11.6 for turbulent rough flow and much smaller than 11.6 for turbulent smooth flow. A composite
form that represents both ranges, as well as the transitional range between them, can be written as

(6.8)

with Bs as a function of Re * = u*ks/v, which can be estimated with

(6.9)

as proposed by Yalin (1992).

6.2.2. Relations for Channel Resistance


Most river flows are indeed hydraulically rough. Equation (6.7) can be used to obtain an approximate expression for depth-
averaged velocity U that is reasonably accurate for many flows. Using the following integral:

(6.10)

but changing the lower limit slightly to avoid the fact that the logarithmic law is singular atz = 0, the following result is obtained:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.11)

or, performing the integration

(6.12)

This relation is known as Keulegan's resistance relation for rough flow.

An approximation to Keulegan's relation is the Manning-Strickler power form

(6.13)

Between Eqs. (6.2) and (6.12), a resistance relation can be found for bed shear stress:

(6.14)

where the friction coefficient Cf is given by

(6.15)

If Eq. (6.13) is used instead of Eq. (6.12), the friction coefficient takes the form

(6.16)

It is useful to show the relationship between the friction coefficient Cf and the roughness parameters in open-channel flow
relations commonly used in practice. Between Eqs. (6.1) and (6.14), the following form of Chezy's law can be derived:

(6.17)

where the Chezy coefficient Cc is given by the relation

(6.18)

A specific evaluation of Chezy's coefficient can be obtained by substitutingEq. (6.15) into Eq. (6.18). It is seen that the
coefficient is not constant but varies as the logarithm of H/ks. A logarithmic dependence is typically a weak one, partially
justifying the common assumption that Chezy's coefficient in Eq. (6.17) is a constant. Substituting Eq. (6.16) into Eqs. (6.17)
and (6.18), Manning's law is obtained:

(6.19)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where Manning's n is given by

(6.20)

The above relation is often called the Manning-Strickler form of Manning's n.

6.2.3. Fixed-Bed and Movable-Bed Roughness


It is clear that to use the above relations for channel flow resistance, a criterion for evaluatingks is necessary. Nikuradse (1933)
proposed the following criterion: Suppose a rough surface is subjected to a flow. The equivalent roughness ks of that surface is
equal to the diameter of sand grains that, when glued uniformly to a completely smooth wall and then subjected to the same
external conditions, yields the same velocity profile. Nikuradse used sand glued to the inside of pipes to conduct this evaluation.
Extending Nikuradse's concept of equivalent grain roughness to the case of rivers and streams, ks can be assumed to be
proportional to a representative sediment size Dx,

(6.21)

Suggested values of αs, which have appeared in the literature, are listed inTable 6.1 (Yen, 1992). Different sizes of sediment
have been suggested for Dx in Eq. (6.21). Statistically, D50 (the grain size for which 50% of the bed material is finer) is most
readily available and meaningful. Physically, a representative size larger than D50 is more meaningful to estimate flow
resistance because of the dominant effect by large sediment particles.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 6.1 Ratio of Nikuradse Equivalent Roughness Size and Sediment Size for Rivers.

Investigator Measure of Sediment Size, D x αs = k s /Dx

Ackers and White (1973) D35 1.23

Strickler (1923) D50 3.3

Keulegan (1938) D50 1

Meyer-Peter and Muller (1948) D50 1

Thompson and Campbell (1979) D50 2.0

Hammond et al. (1984) D50 6.6

Einstein and Barbarossa (1952) D65 1

Irmay (1949) D65 1.5

Engelund and Hansen (1967) D65 2.0

Lane and Carlson (1953) D75 3.2

Gladki (1979) D80 2.5

Leopold et al. (1964) D84 3.9

Limerinos (1970) D84 2.8

Mahmood (1971) D84 5.1

Hey (1979), Bray (1979) D84 3.5

Ikeda (1983) D84 1.5

Colosimo et al. (1986) D84 3~6

Whiting and Dietrich (1990) D84 2.95

Simons and Richardson (1966) D85 1

Kamphuis (1974) D90 2.0

van Rijn (1982) D90 3.0

In flow over a geometrically smooth, fixed boundary, the apparent roughness of the bedks can be computed using Nikuradse's
approach. However, once the transport of bed material has been instigated, the characteristic grain diameter and the thickness
of the viscous sublayer no longer provide the relevant length scales. The characteristic length scale in this situation is the
thickness of the layer where the sediment particles are being transported by the flow, usually referred to as the bedload layer.

Once the bed shear stress τb exceeds the critical shear stress for particle motion τc, the apparent bed roughness ka can be
estimated as follows (Smith and McLean, 1997):

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.22)

where α0 = 26.3, ks is Nikuradse's fixed-bed roughness, and ρs is the bed sediment density. This approach is particularly
suitable for sand bed rivers.

Under intense sediment transport conditions, bedforms, such as dunes, can develop. In this situation, the apparent roughness
also will be influenced by the form drag caused by the presence of bedforms. Nikuradse's approach is valid only for grain-
induced roughness. Methods for flow resistance in the presence of both bedforms and grain roughness are presented later.

6.3. SEDIMENT PROPERTIES


6.3.1. Rock Types
The solid phase of the problem embodied in sediment transport can be any granular substance. In engineering applications,
however, the granular substance in question typically consists of fragments ultimately derived from rocks–hence the name
sediment transport. The properties of these rock-derived fragments, taken singly or in groups of many particles, all play a role in
determining the transportability of the grains under fluid action. The important properties of groups of particles include porosity
and size distribution. The most common rock type one is likely to encounter in the river or coastal environment is quartz. Quartz
is a highly resistant rock and can travel long distances or remain in place for long periods without losing its integrity. Another
highly resistant rock type that is often found together with quartz is feldspar. Other common rock types include limestone,
basalt, granite, and more esoteric types, such as magnetite. Limestone is not a resistant rock; it tends to abrade to silt rather
easily. Silt-sized limestone particles are susceptible to solution unless the water is buffered sufficiently. As a result, limestone
typically is not a major component of sediments at locations distant from its source. On the other hand, it often can be the
dominant rock type in mountain environments.

Basaltic rocks tend to be heavier than most rocks composing the earth’s crust and typically are brought to the surface by
volcanic activity. Basaltic gravels are relatively common in rivers that derive their sediment supply from areas subjected to
vulcanism in recent geologic history. Basaltic sands are much less common. Regions of weathered granite often provide
copious supplies of sediment. Although the particles produced by weathering are often in the granule size, they often break
down quickly to sand size.

Sediments in the fluvial or coastal environment in the size range of silt, or coarser, are generally produced by mechanical
means, including fracture or abrasion. The clay minerals, on the other hand, are produced by chemical action. As a result, they
are fundamentally different from other sediments in many ways. Their ability to absorb water means that the porosity of clay
deposits can vary greatly over time. Clays also display cohesivity, which renders them more resistant to erosion.

6.3.2. Specific Gravity


The specific gravity of sediment is defined as the ratio between the sediment density ρs and the density of water ρ. Some typical
specific gravities for various natural and artificial sediments are listed in Table 6.2.

6.3.3. Size
Herein, the notation D is used to denote sediment size, the typical units of which are millimeters (mm) for sand and coarser
material or microns (μ) for clay and silt. Another standard way of classifying grain sizes is the sedimentological Φ scale,
according to which

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.23)

Table 6.2 Specific Gravity of Rock Types and Artificial Material

Rock type or material Specific gravity ρ s /ρ

quartz 2.60 ~ 2.70

limestone 2.60 ~ 2.80

basalt 2.70 ~ 2.90

magnetite 3.20 ~ 3.50

plastic 1.00 ~ 1.50

coal 1.30 ~ 1.50

walnut shells 1.30 ~ 1.40

Taking the logarithm of both sides, it is seen that

(6.24)

Note that the size Φ = 0 corresponds to D = 1 mm. The usefulness of the Φ scale will become apparent upon a consideration of
grain size distributions. The minus sign has been inserted in Eq. (6.24) simply as a matter of convenience to sedimentologists,
who are more accustomed to working with material finer than 1 mm than they are with coarser material. The reader should
always recall that larger Φ implies finer material. The Φ scale provides a simple way of classifying grain sizes into the following
size ranges in descending order: boulders, cobbles, gravel, sand, silt, and clay. (Table 6.3).

Note that the definition of clay according to size (D < 2μ) does not always correspond to the definition of clay according to
mineral. That is, some clay-mineral particles can be coarser than this limit, and some silt-sized particles produced by grinding
can be finer than that. In general, however, the effect of viscosity makes it difficult to grind up particles in water to sizes finer
than 2μ.

In practical terms, there are several ways to determine grain size. The most popular way for grains ranging from Φ = −4 to Φ = –
4 (0.0625 to 16 mm) is with the use of sieves. Each sieve has a square mesh, the gap size of which corresponds to the
diameter of the largest sphere that would fit through it. Thus, the grain size D so measured corresponds exactly to the diameter
only in the case of a sphere. In general, the sieve size D corresponds to the smallest sieve gap size through which a given grain
can be fitted.

For coarser grain sizes, it is customary to approximate the grain as an ellipsoid. Three lengths can be defined. The length along
the major (longest) axis is denoted as a, the length along the intermediate axis is denoted as b, and the length along the minor
(smallest) axis is denoted as c. These lengths are typically measured with a caliper. The value b is then equated to grain size D.

For grains in the silt and clay sizes, many methods (hydrometer, sedigraph, and so forth) are based on the concept of equivalent
fall diameter. That is, the terminal fall velocity vs of a grain in water at a standard temperature is measured. The equivalent fall
diameter D is the diameter of the sphere having exactly the same fall velocity under the same conditions. Sediment fall velocity
is discussed in more detail below.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
A variety of other more recent methods for sizing fine particles rely on blockage of light beams. The blocked area can be used
to determine the diameter of the equivalent circle: i.e., the projection of the equivalent sphere. It can be seen that all the above
methods can be expected to operate consistently as long as grains shape does not deviate too greatly from a sphere. In
general, this turns out to be the case. There are some important exceptions, however. At the fine end of the spectrum, mica
particles tend to be platelike; the same is true of shale grains at the coarser end. Comparison with a sphere is not necessarily an
especially useful way to characterize grain size for such materials.

6.3.4. Size Distribution


Any sample of sediment normally contains a range of sizes. An appropriate way to characterize these samples is by grain size
distribution. Consider a large bulk sample of sediment of given weight. Let pf(D)—or pf(Φ)—denote the fraction by weight of
material in the sample of material finer than size D(Φ). The customary engineering representation of the grain size distribution
consists of a plot of pf ·100 (percentage finer) versus log10(D): that is, a semilogarithmic plot is used. The same size distribution
plotted in sedimentological form would involve plotting pf·100 versus Φ on a linear plot.

Table 6.3 Sediment Grade Scale

Class Name Size Range Approximate Sieve Mesh Openings per Inch

Millimeters Φ Microns Inches Tyler U.S. standard

SOURCE: Adapted from Vanoni, 1975.

Very large boulders 4,096 ~ 2,048 160 ~ 80

Large boulders 2,048 ~ 1,024 80 ~ 40

Medium boulders 1,024 ~ 512 40 ~ 20

Small boulders 512 ~ 256 −9 ~ −8 20 ~ 10

Large cobbles 256 ~ 128 −8 ~ −7 10 ~ 5

Small cobbles 128 ~ 64 −7 ~ −6 5 ~ 2.5

Very coarse gravel 64 ~ 32 −6 ~ −5 2.5 ~ 1.3

Coarse gravel 32 ~ 16 −5 ~ −4 1.3 ~ 0.6

Medium gravel 16 ~ 8 −4 ~ −3 0.6 ~ 0.3 2 ~ 1/2

Fine gravel 8~4 −3 ~ −2 0.3 ~ 0.16 5 5

Very fine gravel 4~2 −2 ~ −1 0.16 ~ 0.08 9 10

Very coarse sand 2.000 ~ 1.000 −1 ~ 0 2,000 ~ 1,000 16 18

Coarse sand 1.000 ~ 0.500 0~1 1,000 ~ 500 32 35

Medium sand 0.500 ~ 0.250 1~2 500 ~ 250 60 60

Fine sand 0.250 ~ 0.125 2~3 250 ~ 125 115 120

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Class Name Size Range Approximate Sieve Mesh Openings per Inch

Millimeters Φ Microns Inches Tyler U.S. standard

Very fine sand 0.125 ~ 0.062 3~4 125 ~ 62 250 230

Coarse silt 0.062 ~ 0.031 4~5 62 ~ 31

Medium silt 0.031 ~ 0.016 5~6 31 ~ 16

Fine silt 0.016 ~ 0.008 6~7 16 ~ 8

Very fine silt 0.008 ~ 0.004 7~8 8~4

Coarse clay 0.004 ~ 0.0020 8~9 4~2

Medium clay 0.0020 ~ 0.0010 2~1

Fine clay 0.0010 ~ 0.0005 1 ~ 0.5

Very fine clay 0.0005 ~ 0.00024 0.5 ~ 0.24

The size distribution pf(Φ) and size density p(Φ) by weight can be used to extract useful statistics concerning the sediment in
question. Let x denote some percentage, say 50%; the grain size Φx denotes the size such that x percent of the weight of the
sample is composed of finer grains. That is, Φx is defined such that

(6.25)

It follows that the corresponding grain size of equivalent diameter is given byDx, where

(6.26)

The most commonly used grain sizes of this type are the median size D50 and the size D90: i.e., 90% of the sample by weight
consists of finer grains. The latter size is especially useful for characterizing bed roughness.

The density p(Φ) can be used to extract statistical moments. Of these, the most useful are the mean size Φm and the standard
deviation σ. These are given by the relations.

(6.27a, b)

The corresponding geometric mean diameter Dg and geometric standard deviation σg are given as

(6,28a,b)

Note that for a perfectly uniform material, σ = 0 and σg = 1. As a practical matter, a sediment mixture with a value of σg less
than 1.3 is often termed well sorted and can be treated as a uniform material. When the geometric standard deviation exceeds
1.6, the material can be said to be poorly sorted (Diplas and Sutherland, 1988).

In fact, one never has the continuous function p(Φ) with which to compute the moments of Eqs. (6.27a, and b). Instead, one
must rely on a discretization. To this end, the size range covered by a given sample of sediment is discretized using n intervals

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
bounded by n + 1 grain sizes Φ1 , Φ2 ,…, Φn+1 in ascending order of Φ. The following definitions are made from i = 1 to n:

(6.29a)

(6.29b)

Eqs. (6.27a and b) now discretize to

(6.30)

In some cases, especially when the material in question is sand, the size distribution can be approximated as gaussian on the Φ
scale (i.e., log-normal in D). For a perfectly Gaussian distribution, the mean and median sizes coincide:

(6.31)

Furthermore, it can be demonstrated from a standard table of the Gauss distribution that the size Φ displaced one standard
deviation larger that Φm is accurately given by Φ84; by symmetry, the corresponding size that is one standard deviation smaller
than Φm is Φ16. The following relations thus hold:

(6.32a)

(6.32b)

Rearranging the above relations with the aid of Eqs. (6.28a and b) and Eqs. (6.31 and 6.32a),

(6.33a)

(6.33b)

It must be emphasized that the above relations are exact only for a gaussian distribution in Φ. This is not often the case in
nature. As a result, it is strongly recommended that Dg and σg be computed from the full size distribution via Eqs. (6.30a and b)
and (6.28a and b) rather than the approximate form embodied in the above relations.

6.3.5. Porosity
The porosity λp quantifies the fraction of a given volume of sediment that is composed of void space. That is,

If a given mass of sediment of known density is deposited, the volume of the deposit must be computed, assuming that at least
part of it will consist of voids. In the case of well-sorted sand, the porosity often can take values between 0.3 and 0.4. Gravels

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
tend to be more poorly sorted. In this case, finer particles can occupy the spaces between coarser particles, thus reducing the
void ratio to as low as 0.2. Because so-called open-work gravels are essentially devoid of sand and finer material in their
interstices, they may have porosities similar to sand. Freshly deposited clays are notorious for having high porosities. As time
passes, the clay deposit tends to consolidate under its own weight so that porosity slowly decreases.

The issue of porosity becomes of practical importance with regard to salmon spawning grounds in gravel-bed rivers, for
example (Diplas and Parker, 1985). The percentage of sand and silt contained in the sediment is often referred to as the
percentage of fines in the gravel deposit. When this fraction rises above 20 or 26 percent by weight, the deposit is often
rendered unsuitable for spawning. Salmon bury their eggs within the gravel, and a high fines content implies a low porosity and
thus reduced permeability. The flow of groundwater necessary to carry oxygen to the eggs and remove metabolic waste
products is impeded. In addition, newly hatched fry may encounter difficulty in finding enough pore space through which to
emerge to the surface. All the above factors dictate lowered survival rates. Chief causes of elevated fines in gravel rivers
include road building and clear-cutting of timber in the basin.

6.3.6. Shape
Grain shape can be classified in a number of ways. One of these, theZingg classification scheme, is illustrated here (Vanoni,
1975). According to the definitions introduced earlier, a simple way to characterize the shape of an irregular clast (stone) is by
lengths a, b, and c of the major, intermediate, and minor axes, respectively. If the three lengths are equal, the grain can be said to
be close to a sphere in shape. If a and b are equal but c is much larger, the grain should be rodlike. Finally, if c is much smaller
than b, which in turn, is much larger than a, the resulting shape should be bladelike.

6.3.7. Fall Velocity


A fundamental property of sediment particles is their fall velocity. The relation for terminal fall velocity in quiescent fluidvs can
be presented as

(6.34)

where

(6.35a)

(6.35b)

and the functional relation CD = CD(Rp) denotes the drag curve for spheres. This relation is not particularly useful because it is
not explicit in vs; one must compute fall velocity by trial and error. One can use the equation forCD given below

(6.36)

and the definition

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.37)

to obtain an explicit relation for fall velocity in the form ofRf versus Re p. In Fig. 6.2, the ranges for silt, sand, and gravel are
plotted for ν = 0.01 cm2 /s (clear water at 20ºC) and R = 1.65 (quartz). A good summary of relations for terminal fall velocity for
the case of nonspherical (natural) particles can be found in Dietrich (1982), who also proposed the following useful fit:

(6.38)

where b1 = 2.891394, b2 = 0.95296, b3 = 0.056835, b4 = 0.002892, and b5 = 0.000245

6.3.8. Relation Between Size Distribution and Stream Morphology


The study of sediment properties and, in particular, size distribution is most relevant to the context of stream morphology. The
following discussion points out some of the more interesting issues.

Figure 6.2 Sediment fall velocitydiagram

In Fig. 6.3, several size distributions from the sand-bed Kankakee River in Illinois, are shown (Bhowmik et al., 1980). The
characteristic S shape suggests that these distributions might be approximated by a gaussian curve. The median size D50 falls
near 0.3 to 0.4 mm. The distributions are tight, with a near absence of either gravel or silt. For practical purposes, the material
can be approximated as uniform.

In Fig. 6.4, several size distributions pertaining to the gravel-bed Oak Creek in Oregon, are shown (Milhous, 1973). In gravel-bed
streams, the surface layer (“armor” or “pavement”) tends to be coarser than the substrate (identified as “subpavement” in the
figure). Whether the surface or substrate is considered, it is apparent that the distribution ranges over a much wider range of
grain sizes than is the case in Fig. 6.3. More specifically, in the distributions of the sand-bed Kankakee River, Φ varies from
about 0 to about 3, whereas in Oak Creek, Φ varies from about –8 to about 3. In addition, the distribution of Fig. 6.4 is upward-
concave almost everywhere and thus deviates strongly from the gaussian distribution.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.3 Particle size distribution of bed materials in Kankakee River, Illinois. (Bhowmik et al.,
1980)

Figure 6.4 Size distribution of bed material samples in Oak Creek. Oregon. Source: (Milhous,1973)

These two examples provide a window toward generalization. A river can be loosely classified as sand-bed or gravel-bed
according to whether the median size D50 of the surface material or substrate is less than or greater than 2 mm. The size
distributions of sand-bed streams tend to be relatively narrow and also tend to be S shaped. The size distributions of gravel-bed
streams tend to be much broader and to display an upward-concave shape. Of course, there are many exceptions to this
behavior, but it is sufficiently general to warrant emphasis.

More evidence for this behavior is provided in Fig. 6.5. Here, the grain size distributions for a variety of stream reaches have

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
been normalized using the median size D50. Four sand-bed reaches are included with three gravel-bed reaches. All the sand-bed
distributions are S shaped, and all have a lower spread than the gravel-bed distributions. The standard deviation is seen to
increase systematically with increasing D50(White et al., 1973).

The three gravel-bed size distributions differ systematically from the sand-bed distributions in a fashion that accurately reflects
Oak Creek (Fig. 6.4). The standard deviation in all cases is markedly larger than any of the sand-bed distributions, and the
distributions are upward-concave except perhaps near the coarsest sizes.

Figure 6.5 Dimensionless grain-size distribution for different rivers (White et al., 1973)

6.4. THRESHOLD CONDITION FOR SEDIMENT MOVEMENT


When a granular bed is subjected to a turbulent flow, virtually no motion of the grains is observed at some flows, but the bed is
mobilized noticeably at other flows. Factors that affect the mobility of grains subjected to a flow are summarized below:

In the presence of turbulent flow, random fluctuations typically prevent the clear definition of a critical, or threshold condition for
motion: The probability for the movement of a grain is never precisely zero (Lavelle and Mofjeld, 1987). Nevertheless, it is
possible to define a condition below which movement can be neglected for many practical purposes.

6.4.1. Granular Sediment on a Stream Bed


Figure 6.6 is a diagram showing the forces acting on a grain in a bed of other grains. When critical conditions exist and the
grain is on the verge of moving, the moment caused by the critical shear stress τc about the point of support is just equal to that
of the weight of the grain. Equating these moments gives (Vanoni, 1975):

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.39)

Figure 6.6 Forces acting on a sediment particle on an inclined bed

in which γs = specific weight of sediment grains, γ = specific weight of water,D = diameter of grains, is the slope angle of the
stream, θ = the angle of repose of the sediment, c1 and c2 are dimensionless constants, and a1 and a2 are lengths shown in Fig.
6.6. Any consistent set of units can be used inEq. (6.39). For a horizontal bed, Eq. (6.39) reduces to

(6.40)

For an adverse slope (i.e., < 0),

(6.41)

Equations (6.39), (6.40), and (6.41) cannot be used to give τc because the factors c1 , c2 , a1 , and a2 are not known. Therefore,
the relation between the pertinent quantities is expressed by dimensional analysis, and the actual relation is determined from
experimental data. Figure 6.7 is such a relation, first presented by Shields (1936) and carries his name. The curve is expressed
by dimensionless combinations of critical shear stress τc, sediment and water specific weights γs and γ, sediment size D,
critical shear velocity u*c and kinematic viscosity of water v.

These quantities can be expressed in any consistent set of units. Dimensional analysis yields,

(6.42)

The Shields values of are commonly used to denote conditions under which bed sediments are stable but on the verge of
being entrained. Not all workers agree with the results given by the Shields curve. For example, some workers give = 0.047
for the dimensionless critical shear stress for values of R* = u*D/v in excess of 500 instead of 0.06, as shown in Fig. 6.7. Taylor
and Vanoni (1972) reported that small but finite amounts of sediment were transported in flows with values of given by the
Shields curve.

The value of τ c to be used in design depends on the particular case at hand. If the situation is such that grains that are moved
can be replaced by others moving from upstream, some motion can be tolerated, and the Shields values can be used. On the
other hand, if grains removed cannot be replaced, as on a stream bank, the Shields value of τc are too large and should be
reduced.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The Shields diagram is not especially useful in the form of Fig. 6.7 because to find τc, one must know The relation
can be cast in explicit form by plotting versus Re p, noting the internal relation

(6.43)

where is the submerged specific gravity of the sediment. A useful fit is given by Brownlie (1981a):

(6.44)

Figure 6.7 Shields diagram for initiation of motion. Source Vanoni (1975)

With this relation, the value of can be computed readily when the properties of the water and the sediment are given.

The value of bed-shear stress τb for a wide rectangular channel is given by τb = γHS, as shown earlier. The average bed-shear
stress for any channel is given by τb = γRhS, in which Rh = the hydraulic radius of the channel cross section.

6.4.2. Granular Sediment on a Bank


A sediment grain on a bank is less stable than one on the bed because the gravity force tends to move it downward (Ikeda,
1982). The ratio of the critical shear stress τwc for a particle on a bank to that for the same particle on the bed τc is (Lane, 1955)

(6.45)

where is the slope of the bank and θ is the angle of repose for the sediment. Values of θ are given inFig. 6.8 after Lane
(1955) and also can be found in Simons and Senturk (1976).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.8 Angle of repose of granular material. (Lane, 1955)

6.4.3. Granular Sediment on a Sloping Bed


Equation (6.39) shows that τc diminishes as the slope angle increases. For extremely small τc is given by Eq. (6.40).
Taking the ratio between Eqs. (6.39) and (6.40) yields

(6.46)

is the critical shear stress for sediment on a bed with a slope angle and τco is the critical shear stress for a bed with an
extremely small slope. The value of τco can be found from the Shields diagram or with Eq. (6.44). Equation (6.46) is for positive
which is positive for downward sloping beds. For beds with adverse slope, is negative and the term tan in Eq.
(6.46) is positive.

6.4.4. Sediment Mixtures


Several authors have offered empirical or quasi-theoretical extensions of the above relations to the case of mixtures (e.g.,
Wilcock, 1988). Let Di denote the characteristic grain size of the ith size range in a mixture. Furthermore, letDsg denote the
geometric mean size of the surface (exchange, active) layer. Most of the generalizations can be written in the following form
(Parker, 1990):

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.47)

Here

(6.48a)

and

(6.48b)

where τbci and τbcsg denote the values of the dimensioned critical shear stress required to move sediment of sizes Di and Dsg in
the mixture, respectively, and β is an exponent taking a value given below;

(6.49)

Figure 6.9 shows the similarity between four different published expressions having the general form given byEq. (6.47), which
is of interest because it includes the effect of hiding. For uniform material, the critical Shields stress is defined by Eq. (6.44).
Consider two flumes, one with uniform size Da and the other with uniform size Db. For sufficiently coarse material (u*D/v » 1 or
Re p » 1), the critical Shields stress must be the same for both sizes F
( ig. 6.7). It follows from Eq. (6.42) that where τbca and τbcb
denote the dimensioned boundary shear stresses for cases a and b respectively,

(6.50)

Figure 6.9 Critical shear stress for sediment mixture (Source: Misri et al., 1983)

For the case of mixtures, on the other hand, it is seen from Eqs.(6.47) and (6.48) that

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.51)

Comparing Eqs. (6.50) and (6.51), it is seen that a finer particle (Db < Da, or alternatively, Di < Dsg ) is more mobile than a coarser
particle. For example, suppose that one grain size is four times coarser than another. If two uniform sediments are being
compared, it follows from Eq. (6.50) that the critical shear stress for the coarser material is four times that of the finer material.
In the case of a mixture, however, the critical shear stress for the coarser material is only about 40.1, or 1.15 times that for the
finer material.

A finer particle in a mixture is thus seen to be only a little more mobile than its coarser-sized brethren, where uniform beds of
fine material are much more mobile than are uniform beds of coarser material. The reason is that finer particles in a mixture
are relatively less exposed to the flow; they tend to hide in the lee of coarser particles. By the same token, a particle is relatively
more exposed to the flow when most of its neighbors are finer.

A method to calculate the critical shear stress for motion of uniform and heterogeneous sediments was proposed by Wiberg
and Smith (1987) on the basis of the fluid mechanics of initiation of motion, which takes into account both roughness and
hiding effects.

6.5. SEDIMENT TRANSPORT


6.5.1. Sediment Transport Modes
The most common modes of sediment transport in rivers are bedload and suspended load. In the case of bedload, the particles
roll, slide, or saltate over each other, never deviating too far above the bed. In the case of suspended load, the fluid turbulence
comes into play carrying the particles well up into the water column. In both cases, the driving force for sediment transport is
the action of gravity on the fluid phase; this force is transmitted to the particles via drag.

The same phenomena of bedload and suspended load transport occur in a variety of other geophysical contexts. Sediment
transport is accomplished in the near-shore lake and oceanic environment by wave action. Turbidity currents carry sediment
into lakes, reservoirs, and the deep sea.

The phenomenon of sediment transport can sometimes be disguised in rather esoteric phenomena. When water is
supercooled, large quantities of particulate frazil ice can form. As this water moves under a frozen ice cover, the phenomenon
of sediment transport in rivers is stood on its head. The frazil ice particles float rather than sink and thus tend to accumulate on
the bottom side of the ice cover rather than on the river bed. Turbulence tends to suspend the particles downward rather than
upward.

In the case of a powder snow avalanche, the fluid phase is air and the solid phase consists of snow particles. The dominant
mode of transport is suspension. These flows are close analogies of turbidity currents, insofar as the driving force for the flow
is the action of gravity on the solid phase rather than the fluid phase. That is, if all the particles drop out of suspension, the flow
ceases. In the case of sediment transport in rivers, it is accurate to say that the fluid phase drags the solid phase along. In the
case of turbidity currents and powder snow avalanches, the solid phase drags the fluid phase along.

Desert sand dunes provide an example for which the fluid phase is air, but the dominant mode of transport is saltation rather
than suspension. Because air is so much lighter than water, quartz sand particles saltate in long, high trajectories, relatively
unaffected by the direct action of turbulent fluctuations. The dunes themselves are created by the effect of the fluid phase
acting on the solid phase. They, in turn, affect the fluid phase by changing the resistance.

Among the most interesting sediment–transport phenomena are debris flows, slurries, and hyperconcentrated flows. In all
these cases, the solid and fluid phases are present in similar quantities. A debris flow typically carries a heterogeneous mixture

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
of grain sizes ranging from boulders to clay. Slurries and hyperconcentrated flows are generally restricted to finer grain sizes. In
most cases, it is useful to think of such flows as consisting of a single phase, the mechanics of which are highly non-
Newtonian.

The study of the movement of grains under the influence of fluid drag and gravity becomes even more interesting when one
considers the link between sediment transport and morphology. In the laboratory, the phenomenon can be studied in the
context of a variety of containers, such as channel and wave tanks, specified by the experimentalist. In the field, however, the
fluid-sediment mixture constructs its own container. This new degree of freedom opens up a variety of intriguing possibilities.

Consider the river. Depending on the existence or lack of a viscous sublayer and the relative importance of bedload versus
suspended load, a variety of rhythmic structures can form on the river bed. These include ripples, dunes, antidunes, and
alternate bars. The first three of these can have a profound effect on the resistance to flow offered by the river bed. Thus, they
act to control river depth. River banks themselves also can be considered to be a self-formed morphological feature, thus
specifying the entire container.

The container itself can deform in plan. Alternate bars cause rivers to erode their banks in a rhythmic pattern, thus allowing for
the onset of meandering. Fully developed river meandering implies an intricate balance between sediment erosion and
deposition. If a stream is sufficiently wide, it will braid rather than meander, dividing into several intertwining channels.

Rivers create morphological structures at much larger scales as well. These include canyons, alluvial fans, and deltas. Turbidity
currents create similar structures in the oceanic environment. In the coastal environment, the beach profile itself is created by
the interaction of water and sediment. On a larger scale, offshore bars, spits, and capes constitute rhythmic features created by
wave-current-sediment interaction. The boulder levees often created by debris flows provide another example of a morphologic
structure created by a sediment-bearing flow.

The floodplains of most sand-bed rivers often contain copious amounts of silt and clay finer than approximately 50 μ. This
material is often called wash load because it moves through the river system without being present in the bed in significant
quantities. Increased wash load does not cause deposition on the bed, and reduced wash load does not cause erosion because
it is transported well below capacity. This is not meant to imply that the wash load does not interact with the river system. Wash
load in the water column exchanges with the banks and the floodplain rather than the bed. Greatly increased wash load, for
example, can lead to thickened floodplain deposits, with a consequent increase in bankfull channel depth.

The emphasis here is the understanding of bedload and suspended load transport in rivers, with the goal of providing the
knowledge needed to do sound sedimentation engineering, particularly with problems involving stream restoration and
naturalization.

6.5.2. Shields Regime Diagram


In the context of rivers, it is useful to have a way to determine what kind of sediment-transport phenomena can be expected for
different flow conditions and different characteristics of sediment particles. In Fig. 6.10, the ordinates correspond to bed shear
stresses written in the dimensionless form proposed by Shields

(6.52)

and the particle Re p, defined by Eq. (6.37) is used for the abscissa values. There are three curves in the diagram which make it
possible to know, for different values of (τ*, Re p), if the given bed sediment will go into motion, and if this is the case whether or
not the prevailing mode of transport will be in suspension or as bedload. The diagram also can be used to predict what kind of
bedforms can be expected. For example, ripples will develop in the presence of a viscous sublayer and fine-grained sediment. If
the viscous sublayer is disrupted by coarse sediment particles, then dunes will be the most common type of bedform.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The Shields regime diagram also shows a clear distinction between the conditions observed in sand-bed rivers and gravel-bed
rivers at bankfull stage. If one wanted to model in the laboratory sediment transport in rivers, the experimental conditions
would be different, depending on the river system in question. As could be expected, the diagram also shows that in gravel-bed
rivers, sediment is transported as bedload. In sand-bed rivers, on the other hand, suspended load and bedload transport coexist
most of the time.

The regime diagram is valid for steady, uniform, turbulent flow conditions, where the bed shear stress τb can be estimated with
Eq. (6.1). The ranges for silt, sand, and gravel also are included. In the diagram, the critical Shields stress for motion was
plotted with the help of Eq. (6.44). The critical condition for suspension is given by the following ratio:

(6.53)

Figure 6.10 Shields regime diagram. (Source: Gary Parker)

where u* is the shear velocity and vs is the sediment fall velocity. Equation (6.53) can be transformed into

(6.54)

where

(6.55)

and Rf is given by Eq. (6.35a) and can be computed for different values of Re p with the help of Eq. (6.38).

Finally, the critical condition for viscous effects (ripples) was obtained with the help of Eq. (6.5) as follows:

(6.56)

which in dimensionless form can be written as

(6.57)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Relations (6.44), (6.54), and (6.57) are the ones plotted in Fig. 6.10. The Shields regime diagram should be useful for studies
concerning stream restoration and naturalization because it provides the range of dimensionless shear stresses corresponding
to bankfull flow conditions for both gravel- and sand-bed streams.

6.6. BEDLOAD TRANSPORT


6.6.1. The Bed Load Transport Function
Bedload particles roll, slide, or saltate along the bed. The transport thus occurs tangential to the bed. In a case where all the
transport is directed in the streamwise, or s direction, the volume bedload-transport rate per unit width (n direction) is given by
q; the units are length3 /length/per time, or length2 /time. In general, q is a function of boundary shear stress τb and other
parameters; that is,

(6.58)

In general, bedload transport is vectorial, with components qs and qn in the s and n directions, respectively.

6.6.2. Erosion Into and Deposition from Suspension


The volume rate of erosion of bed material into suspension per unit time per unit bed area is denoted asE. The units of E are
length3 /length2 /time, or velocity. A dimensionless sediment entrainment rate Es can thus be defined with the sediment fall
velocity vs:

(6.59)

In general, Es can be expected to be a function of boundary shear stress τb and other parameters. Erosion into suspension can
be taken to be directed upward normal: i.e., in the positive z direction.

Let denote the volume concentration of suspended sediment (m3 of sediment/m3 of sediment-water mixture), averaged over
turbulence. The streamwise volume transport rate of suspended sediment per unit width is given by

(6.60)

In a two-dimensional case, two components, qSs and qSn, result, where

(6.61a)

(6.61b)

Deposition onto the bed is by means of settling. The rate at which material is fluxed vertically downward onto the bed
(volume/area/time) is given by where is a near-bed value of The deposition rate D realized at the bed is obtained by
computing the component of this flux that is actually directed normal to the bed:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.62)

6.6.3. The Exner Equation of Sediment Mass Conservation for Uniform


Material
Consider a portion of river bottom, where the bed material is taken to have a (constant) porosity λ.
p Mass balance of sediment
requires the following equation to be satisfied:

A datum of constant elevation is located well below the bed level, and the elevation of the bed with respect to such datum is
given by η. Then, bed level changes as a result of bedload transport, sediment entrainment into suspension, and sediment
deposition onto the bed can be predicted with the help of

(6.63)

To solve the Exner equation, it is necessary to have relations to compute bedload transport (i.e.,qs and qn), near-bed suspended
sediment concentration and sediment entrainment into suspension Es. The basic form of Eq. (6.63) was first proposed by
Exner (1925).

6.6.4. Bedload Transport Relations


A large number of bedload relations can be expressed in the general form

(6.64)

Here, q* is a dimensionless bedload transport rate known as the Einstein number, first introduced by H. A. Einstein in 1950 and
given by

(6.65)

The following relations are of interest. In 1972, Ashida and Michiue introduced

(6.66)

and recommend a value of of 0.05. It has been verified with uniform material ranging in size from 0.3 mm to 7 mm. Meyer-
Peter and Muller (1948) introduced the following:

(6.67)

where This formula is empirical in nature and has been verified with data for uniform gravel.

Engelund and Fredsøe (1976) proposed,

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.68)

where This formula resembles that of Ashida and Michiue because the derivation is almost identical.

Fernandez Luque and van Beek (1976) developed the following,

(6.69)

where varies from 0.05 for 0.9 mm material to 0.058 for 3.3. mm material. The relation is empirical in nature.

Wilson (1966):

(6.70)

where was determined from the Shields diagram. This relation is empirical in nature; most of the data used to fit it pertain to
very high rates of bedload transport.

Einstein (1950):

(6.71)

where the functionality is implicitly defined by the relation

(6.72)

Note that this relation contains no critical stress. It has been used for uniform sand and gravel.

Yalin (1963):

(6.73)

where

(6.74)

and is evaluated from a standard Shields curve. Two constants in this formula have been evaluated with the aid of data
quoted by Einstein (1950), pertaining to 0.8 mm and 28.6 mm material.

Parker (1978):

(6.75)

developed with data sets pertaining to rough mobile-bed flow over gravel.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Several of these relations are plotted in Fig. 6.11. They tend to be rather similar in nature. Scores of similar relations could be
quoted.

To date, only few research groups have attempted complete derivations of the bedload function in water. They are Wiberg and
Smith (1989), Sekine and Kikkawa (1992), García and Niño (1992), Niño and García, (1994, 1998), and Niño et al., (1994).

Figure 6.11 Bedload transport relations. (Parker, 1990)

6.6.5. Bedload Transport Relation for Mixtures.


Relatively few bedload relations have been developed specifically in the context of mixtures (e.g., Bridge and Bennett, 1992).
One of these is presented below as an example.

The relationship of Parker (1990) applies to gravel-bed streams. The data used to fit the relation are solely from two natural
gravel-bed streams: Oak Creek in Oregon and the Elbow River in Alberta, Canada. The relation is surface-based; load is specified
per unit of fractional content in the surface layer. The surface layer is divided into N size ranges, each with a fractional content
Fi by volume, and a mean phi size The arithmetic mean of the surface size on the phi scale and the
corresponding arithmetic standard deviation are given by

(6.76a, b)

The corresponding geometric mean size Dsg and the geometric standard deviation σsg of the surface layer are given by

(6.77a, b)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
In the Parker relation, the volume bedload transport per unit width of gravel in theith size range is given by the product qiFi (no
summation), where qi denotes the transport per unit fraction in the surface layer. The total volume bedload transport rate of
gravel per unit width is qT, where

(6.78)

The relation does not apply to sand. Thus, before using the relation for a given surface distribution, the sand content of the
grain-size distribution must be removed and Fi must be renormalized so that it sums to unity over all sizes in excess of 2 mm.

If pi denotes the fraction volume content of material in the ith size range in the bedload, it follows that

(6.79)

The parameter qi is made dimensionless as follows:

(6.80)

A dimensionless Shields stress based on the surface geometric mean size is defined as follows:

(6.81)

Let denote a normalized value of this Shields stress, given by

(6.82)

where

(6.83)

corresponds to a “near-critical” value of Shields stress. The Parker relation can then be expressed in the form

(6.84a)

In the above relationship, go denotes a hiding function given by

(6.84b)

The parameter ω is given by the relationship

(6.84c)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where and ω o are specified as functions of in Fig. 6.12. The function G is specified as

(6.85)

and is shown in Fig. 6.13. Here, Mo = 14.2 and is a dummy variable for the argument in Eq. (6.84) and is not to be confused
with the grain-size scale.

An application of Eq. (6.84) to uniform material with size D results in the relation

(6.86)

where

(6.87)

and q denotes the volumetric sediment transport per unit width. In Fig. 6.11, Eq. (6.86) is compared to several other relations
and selected laboratory data for uniform material. The figure is adapted from Figs. 6b and 7 in Wiberg and Smith (1989), where
reference to the data and equations can be found. The data pertain to 0.5 mm sand and 28.6 mm gravel. Equation (6.86) shows
a reasonable correspondence with the data and with several other relations for uniform material.

The Parker relationship (Eq. 6.84) can be used to predict mobile or static armor in gravel streams. Note that there is no formal
critical stress in the formulation; instead for ≤ 1, the transport rates become extremely small. For the computation of bedload
transport in poorly sorted gravel-bed rivers, the above formulation has been used to implement a series of programs named
“ACRONYM” (Parker, 1990). The program “ACRONYM1” provides an implementation of the surface-based bedload transport
equation presented in Parker (1990). It computes the magnitude and size distribution of bedload transport over a bed surface
of given size distribution, on which a given boundary shear stress is imposed. The program “ACRONYM2” inverts the same
bedload transport equation, allowing for calculation of the size distribution at a given boundary shear stress. The program was
used to compute mobile and static armor size distributions in Parker (1990) and Parker and Sutherland (1990).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.12 Plots of ω0 and σϕ0 versus ϕsg0 the asymptotes are noted on the plot. (Parker, 1990)

Figure 6.13 Plot of G and GT versus ϕ50. (Parker, 1990)

The program “ACRONYM3” allows for the computation of aggradation or degradation to a specified active or static equilibrium
final state. To this end, Parker’s method (1990) is combined with a resistance relation of the Keulegan type. In the program,
both constant width and water discharge are assumed.

The program “ACRONYM4” is directed toward the wavelike aggravation of self-similar form discussed in Parker (1991a, 1991b).
It uses Parker’s method and a resistance relation of the Manning-Strickler type to compute downstream fining and slope
concavity caused by selective sorting and abrasion.

6.7. BEDFORMS
The formation and behavior of sediment waves produced by moving water are, in equal measure, intellectually intriguing and of
great engineering importance. Because of the central role they play in river hydraulics, fluvial ripples, dunes, and bars have

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
received extensive attention from engineers for at least the past two centuries, and even more intensive descriptive study from
geologists. Such studies can be divided into three categories according to the approach followed: analytical, empirical, or
statistical.

Analytical models for bedforms have been proposed since 1925 (Anderson, 1953; Blondeaux et al., 1985; Colombini et al., 1987;
Engelund, 1970; Exner, 1925; Fredsoe, 1974, 1982; Gill, 1971; Haque and Mahmood, 1985; Hayashi, 1970; Kennedy, 1963, 1969;
Parker, 1975; Raudkivi and Witte, 1990; Richards, 1980; Smith, 1970; Tubino and Seminara, 1990).

Empirical methods include the following works (Coleman and Melville, 1994; Colombini et al., 1990; García and Niño, 1993;
Garde and Albertson, 1959; Ikeda, 1984; Jaeggi, 1984; Kinoshita and Miwa, 1974; Menduni and Paris, 1986: Ranga Raju and
Soni, 1976; Raudkivi, 1963; van Rijn, 1984; Yalin, 1964; Yalin and Karahan, 1979).

Statistical models for bedforms have been advanced by the following authors Annambhotla et al.,1972; Hino, 1968; Jain and
Kennedy, 1974; Nakagawa and Tsujimoto, 1984; Nordin and Algert, (1966).

Despite all the research that has been done, there is presently no completely reliable predictor for the conditions of occurrence
and characteristics of the different bed configurations (ripples, dunes, flat bed, antidunes).

6.7.1. Dunes, Antidunes, Ripples, and Bars


The ripples, dunes, and antidunes illustrated in Fig. 6.14 are the classic bedforms of erodible-bed open-channel flow. On the one
hand, they are a product of the flow and sediment transport; on the other hand, they profoundly influence the flow and sediment
transport. In fact, all the bedload formulas quoted previously are strictly invalid in the presence of bedforms. The adjustments
necessary to render them valid are discussed later.

Ripples, dunes, and antidunes are undular (wavelike) features that have wavelengths Λ and wave heights Δ that scale no larger
than on the order of the flow depth H, as defined below.

6.7.1.1. Dunes.
Well-developed dunes tend to have wave heights D scaling up to about one-sixth of the depth: i.e.,

(6.88)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.14 Schematic of different bedforms. (Vanoni, 1975)

Dune wavelengths can vary considerably. A fairly typical range can be quantified as dimensionless wave number k, where

(6.89)

This range is

(6.90)

Dunes invariably migrate downstream. Typically, they are approximately triangular in shape and usually (but not always)
possess a slip face, beyond which the flow is separated for a certain length. A dune progresses forward as bedload accretes on
the slip face. Generally, little bedload is able to pass beyond the face without depositing on it, whereas most of the suspended
load is not directly affected by it.

Let c denote the wave speed of the dune. The bedload transport rate can be estimated as the volume of material transported
forward per unit bed area per unit time by a migrating dune. If the dune is approximated as triangular in shape, the following
approximation holds:

(6.91)

Dunes are characteristic of subcritical flow in the Froude sense. In a shallow-water (long wave) model, the Froude criterion (Fr)
dividing subcritical and supercritical flow is

(6.92)

where

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.93)

Dunes, however, do not qualify as long waves because their wavelength is of the order of the depth. A detailed potential flow
analysis over a wavy bed yields the following (wave-number dependent) criterion for critical flow over a bedform (Kennedy,
1963).

(6.94)

Note that as k → 0(Λ→ ∞) tanh(k) → k, and condition (6.92) is recovered in the long-wave limit. For dunes to occur, then, the
following condition must be satisfied:

(6.95)

Both dunes and antidunes cause the water surface as well as the bed to undulate. In the case of dunes, the undulation of the
water surface is usually of much smaller amplitude than that of the bed; the two are nearly 180o out of phase.

Dunes also can occur in the case of wind-blown sand. Barchan dunes are commonly observed in the desert. In addition, they
can be found in the fluvial environment in the case of sand (in supply insufficient to cover the bed completely) migrating over an
immobile gravel bed.

6.7.1.2. Antidunes.
Antidunes are distinguished from dunes by the fact that the undulations of the water surface are nearly in phase with those of
the bed. They are associated with supercritical flow in the sense that

(6.96)

Antidunes may migrate either upstream or downstream. Upstream-migrating antidunes are usually rather symmetrical in shape
and lack a slip face. Downstream-migrating antidunes are rarer; they have a well-defined slip face and look rather like dunes.
The distinguishing feature is the water surface undulations, which are pronounced in the case of antidunes.

The potential-flow criterion dividing upstream-migrating antidunes from downstream-migrating antidunes is

(6.97)

Values lower than the value in Eq. 6.97 are associated with upstream-migrating antidunes.

6.7.1.3. Ripples.
Ripples are dunelike features that occur only in the presence of a viscous sublayer. They look much like dunes because they
migrate downstream and have a pronounced slip face. They generally are much more three-dimensional in structure than are
dunes, however, and have little effect on the water surface.

A criterion for the existence of ripples is the existence of a viscous sublayer. Recalling that the thickness of the viscous sublayer
is given by δv = 11.6v/u*, it follows that ripples form when

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.98)

6.7.1.4. Bars.
Bars are bedforms in rivers that scale the channel width. They include alternate bars in straight streams, point bars in
meandering streams, and pool bars in braided streams. In straight streams, the minimum channel slope S necessary for
alternate-bar formation is given by

(6.99)

(Jaeggi, 1984), where B is the channel width, Dg is the geometric mean size of the bed sediment, as given byEq. (6.82a), and M
is a parameter that varies from 0.34 for uniformsized bed material to 0.7 for poorly sorted material.

Scour depth (S d ) caused by alternate bar formation can be estimated with

(6.100)

where ΔAB is the total height of the alternate bar.

6.7.1.5. Progression of bedforms.


Various bedforms are associated with various flow regimes. In the case of a sand-bed stream with a characteristic size less
than about 0.5 mm, a clear progression is evident as flow velocity increases. This is illustrated in Fig. 6.14.

The bed is assumed to be initially flat. At low imposed velocityU, the bed remains flat because no sediment is moved. As the
velocity exceeds the critical value, ripples are formed first. At higher values, dunes form and coexist with ripples. For even
higher velocities, well-developed dunes form in the absences of ripples.

At some point, the velocity reaches a value near the critical value in the Froude sense: i.e.,Eq. (6.94). Near this point, the dunes
often are suddenly and dramatically washed out. This results in a flat bed known as an upper-regime (supercritical) flat bed.
Further increases in velocity lead to the formation of antidunes and, finally, to the chute and pool pattern. The last of these is
characterized by a series of hydraulic jumps.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 6.4 Summary of Bedform Effects on Flow Configuration

Bed Form or Dimensions Shape Behavior and Occurrence


Configuration
(2) (3) (4)
(1)

*Reported by Kennedy (1969).

Source: Vanoni (1975).

Ripples Wavelength less Roughly triangular in profile, with gentle, slightly Move downstream with velocity much less than
than approx 1 ft; convex upstream slopes and downstream slopes that of the flow. Generally do not occur in
height less than nearly equal to the angle of repose. Generally sediments coarser than about 0.6 mm.
approx 0.1 ft. short-crested and three-dimensional.

Bars Lengths comparable Profile similar to ripples. Plan form variable. Four types of bars are distinguished: (1) point,
to the channel width. (2) alternating, (3) transverse, and (4) tributary.
Height comparable Ripples may occur on upstream slopes.
to mean flow depth.

Dunes Wavelength and Similar to ripples. Upstream slopes of dunes may be covered with
height greater than ripples. Dunes migrate downstream in manner
ripples but less than similar to ripples.
bars.

Transition Vary widely Vary widely. A configuration consisting of a heterogeneous


array of bed forms, primarily low amplitude
ripples and dunes interspersed with flat
regions.

Flat bed — — A bed surface devoid of bed forms. May not


occur for some ranges of depth and sand size.

Antidunes Wave length = Nearly sinusoidal in profile. Crest length In phase with and strongly interact with gravity
2πV2/g (approx)* comparable to wavelength. watersurface waves. May move upstream,
Height depends on downstream, or remain stationary, depending
depth and velocity of on properties of flow and sediment.
flow.

In the case of a bed coarser than 0.5 mm, the ripple regime is replaced by a zone characterized by a lower-regime (subcritical)
flat bed. Above this lies the ranges for dunes, the upper-regime flat bed, and antidunes.

The effect of bedforms on flow resistance is summarized in Table 6.4. As noted earlier for equilibrium flows in wide straight
channels, the relation for bed resistance can be expressed in the form

(6.101)

where Cf denotes a bed-friction coefficient. If the bed were rigid and the flow were rough,Cf would vary only weakly with the
flow, according to the logarithmic law embodied in Eq. (6.12). As a result, the relation between τb and U is approximately
parabolic.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.15 Variations of bed shear stress τb and Darcy-Weisbach friction factor f with mean
velocity U in flow over a fine sand bed. (Raudkivi, 1990)

The effect of bedforms is to increase the bed shear stress to values often well above that associated with the skin friction of a
rough bed alone. In Fig. 6.15, a plot of τb versus U is given for the case of an erodible bed. At extremely low values ofU, the
parabolic law is followed. As ripples, then dunes are formed, the bed shear stress rises to a maximum value. At this maximum
value, the value of Cf is seen to be as much as five times the value without dunes. It is clear that dunes play an important role
regarding bed resistance. The increased resistance results from form drag in the lee of the dune.

As the flow velocity increases further, dune wavelength gradually increases and dune height diminishes, leading to a gradual
reduction in resistance. At some point, the dunes are washed out, and the parabolic law is again satisfied. At even higher
velocities, the form drag associated with antidunes appears; it is usually not as pronounced as that of dunes.

6.7.2. Dimensionless Characterization of Bedform Regime


Based on the above arguments, it is possible to identify at least three parameters governing bedforms at equilibrium flow.
These are Shields stress τ*, shear Reynolds number Re = u*D/v, and Fr. A characteristic feature of sediment transport is the
proliferation of dimensionless parameters. This feature notwithstanding, Parker and Anderson (1977) showed that equilibrium
relations of sediment transport for uniform material in a straight channel can be expressed with just two dimensionless
hydraulic parameters, along with a particle Re (e.g., Re p or Re) and a measure of the denstiy difference (e.g., R).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.16 Bedform predictor proposed by Simons and Richardson (1966).

In the case of bedforms, then, the following classification can be proposed:

(6.102)

Here, any independent pair of hydraulic variables π1 , π 2 applicable to the problem can be specified because any one pair can be
transformed into any other independent pair. For example, the pair τ* and Fr might be used or, alternatively, S and H/D.

One popular discriminator of bedform type is not expressed in dimensionless form at all. It is the diagram proposed by Simons
and Richardson (1966), (Fig. 6.16). In the diagram, regimes for ripples and dunes, transition to the upper-regime plane bed, and
upper-regime plane bed and antidunes are shown. The two hydraulic parameters are abbreviated to a single one, stream power
τbU, and the particle Re is replaced by grain size D. The diagram is applicable only for sand-bed streams of relatively small
scale.

Liu’s discriminator (1957), shown in Fig. 6.17, uses one dimensionless hydraulic parameter u*/vs (a surrogate for τ*) and the
particle Re p. The diagram is of interest because it covers sizes much coarser than those of Simons and Richardson. It is seen
that the various regimes become compressed as grain size increases. In the case of extremely coarse material, the flow must
be supercritical for any motion to occur. As a result, neither ripples or dunes are expected.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.17 Criteria for bedforms proposed by Liu (1957)

Figure 6.18 Bedform classification. (after Chabert and Chauvin, 1963)

In fact, dunes can occur over a limited range in the case of coarse material. This is illustrated inFig. 6.18. The diagram shows
that Re must be less than approximately 10(δv < D) for ripples to form. Recalling that

(6.103)

and using a critical value of τ* of approximately 0.03, it is seen from Eq. (6.101) and the conditions R = 1.65, v = 0.01 cm2 /s that
the condition Re = 10 corresponds to a value of D of approximately 0.6 mm.

For coarser grain sizes, the dune regime is preceded by a fairly wide range consisting of a lower-regime flat bed. Many gravel-

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
bed rivers never leave this lower-regime flat bed region, even at bankfull flow. The diagram in Fig. 6.18 is not suited to the
description of upper-regime flow.

Figure 6.19a Bed-form chart for Rg = 4.5–10 (D50 = 0.12 mm–0.200 mm)

Figure 6.19b Bed-form chart for Rg = 4.5–10 (D50 = 0.12 mm–0.200 mm)

Figure 6.19c Bed-form chart for Rg = 4.5–10 (D50 = 0.15 mm–0.32 mm)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.19d Bed-form chart for Rg = 16–26 (D50 = 0.228 mm–0.45 mm)

Figure 6.19e Bed-form chart for Rg = 24–48 (D50 = 0.4 mm–0.57 mm)

Figure 6.19f Bedform chart. A, B, C, D, E, F (after Vanoni, 1974)

A complete set of diagrams for the case of sand is shown inFig. 6.19a to f, (Vanoni, 1974). The two hydraulic parameters are Fr
and H/D; the particle Re used in the plot is equal to Re p/R, and constant R is set at 1.65. Note how the transition to upper regime
occurs at progressively lower values of Fr for relatively deeper flow (in the sense that H/D becomes large).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.20 Bedform classification (after van Rijin, 1984)

A bedform classification scheme that includes both the lower and the upper regime was proposed by Van Rijn (1984). The
scheme is based on a dimensionless particle diameter D* and the transport-stage parameter T defined, respectively, as

(6.104)

and

(6.105)

where is the bed shear stress caused by skin or grain friction, and is the critical shear stress for motion from the Shields
diagram.

Van Rijn (1984) suggested that ripples form when both D* < 10 and T < 3, as shown in Fig. 6.20. Dunes are present elsewhere
when T < 15, dunes wash out when 15 < T < 25, and upper flow regime starts when T > 25.

In the lower regime, the geometry of bedforms refers to representative dune height Δ and wavelength Δ as a function of the
average flow depth H, median bed particle diameter D50, and other flow parameters such as the transport-stage parameter T,
and the grain shear Reynolds number Re. The bedform height and steepness predictors proposed by van Rijn (1984) are

(6.106)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
and

(6.107)

Figure 6.21a,b Bedform height and steepness (after van Rijn, 1984)

The bedform length obtained from dividing these two equations, Λ = 7.3H, is close to the theoretical value Λ = 2πH, derived by
Yalin (1964). The agreement with laboratory data is good, as shown in Fig. 6.21a and b, but both curves tend to underestimate
the bedform height and steepness of field data (Julien, 1995; Julien and Klaasen, 1995). For instance, lower-regime bedforms
are observed in the Mississippi River at values of T well beyond 25. Large dunes on alluvial rivers often display small dunes
moving along their stoss face (Amsler and García, 1997; Klaasen et al., 1986), resulting in additional form drag that is not
accounted for in relations derived from laboratory observations. What is needed is a predictor for bedforms in large alluvial
rivers based on field observations,

6.7.3. Effect of Bedforms on River Stage


The presence or absence of bedforms on the bed of a river can lead to some curious effects on a river’s stage. According to a
standard Manning-type relation for an nonerodible bed, the following should hold:

(6.108)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.22 Flow velocity versus hydraulic radius for the Rio Grande (after Nordin, 1964)

Here, the channel is assumed to be wide enough to allow the hydraulic radius to be replaced with the depthH. According to Eq.
(6.108), if the energy slope remains relatively constant, depth should increase monotonically with increasing velocity. This
would indeed be the case for a rigid bed. In a sand-bed stream, however, resistance decreases as U increases over a wide range
of conditions.

At equilibrium,

(6.109)

This decrease in resistance implies that depth does not increase as rapidly in U as it would for a rigid-bed open channel. In fact,
as transition to upper regime is approached, the bedforms can be wiped out suddenly, resulting in a dramatic decrease in
resistance. The result can be an actual decrease in depth as velocity increases (Fig. 6.22).

It is often found that the discharge at which the dunes are obliterated is a little below bankfull in sand-bed streams. As a result,
flooding is not as severe as it would be otherwise. The precise point of transition is generally different, depending on whether
the discharge is increasing or decreasing. This can lead to double-valued stage-discharge relations, (Fig. 6.22).

6.8. EFFECT OF BEDFORMS ON FLOW AND SEDIMENT


TRANSPORT
6.8.1. Form Drag and Skin Friction
As was seen in Sec. 6.7.3, bedforms can have a profound influence on the flow resistance and thus on the sediment transport
in an alluvial channel. To characterize the importance of bedforms in this regard, it is of value to consider the forces that
contribute to the drag force on the bed.

Consider, for example, the case of normal flow in a wide rectangular channel. In the presence of bedforms,Eq. (6.1) must be
amended to

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.110)

where is an effective boundary shear stress, where the overbear denotes averaging over the bedforms and can be defined as
the streamwise drag force per unit area, where H now represents the depth averaged over the bedforms.

In most cases of interest, the two major sources of the effective boundary shear stress are skin friction, which is associated
with the shear stresses, and the form drag, which is associated with the pressure. That is,

(6.111)

where τbs is the shear stress caused by skin friction and τbf is the shear stress caused by form drag.

The important thing to realize is that form drag results from a net pressure distribution over an entire bedform. At any given
point along the surface of the bedform, the pressure force acts normal to the body. For this reason, form drag is ineffective in
either moving bedload sediment or entraining sediment into suspension. In the case of dunes in rivers, because the flow usually
separates in the lee of the crest, the form drag is often substantial. The part of the effective shear stress that governs sediment
transport is thus seen to be the skin friction.

To render any of the bedload formulas presented in Sec. 6.6.4 valid in the presence of bedforms, it is necessary to replace the
Shields stress τ* by the Shields stress associated with skin friction only:

(6.112)

The fact that the form drag needs to be excluded to compute sediment transport does not by any means imply that it is
unimportant. It is often the dominant source of boundary resistance and thus plays a crucial role in determining the depth of
flow. This will be considered in more detail below.

6.8.2. Shear Stress Partitions


6.8.2.1. Einstein partition.
Einstein (1950) was among the first to recognize the necessity to distinguish between skin friction and form drag. He proposed
the following simple scheme to partition the two. Equation (6.101) is amended to represent an effective boundary shear stress
averaged over bedforms:

(6.113)

where Cf now represents a resistance coefficient that includes both skin friction and form drag. For a given flow velocity U,
Einstein computed the skin friction as follows:

(6.l14)

where Cfs is the frictional resistance coefficient that would result if bedforms were absent. For example, in the case of rough
turbulent flow, Eq. (6.15) may be used:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.115)

(In fact, Einstein presented a slightly different formula, which allows for turbulent smooth and transitional flow as well.) The
parameter Hs denotes the depth that would result in the absence of bedforms (but with U held constant). This depth is per force
less than H because the resistance is less in the absence of bedforms.

The remaining problem is how to calculate Hs. Einstein restricted his arguments to the case of normal flow. In this case,Eq.
(6.15) holds: that is,

(6.116a)

and

(6.116b)

Now, between Eqs. (6.113) and (6.116b), the following relation is obtained for Hs:

(6.117)

For given values of U, ks, and S (averaged over bedforms), Eq. (6.117) is easily solved iteratively for Hs. Once Hs is known, it is
not difficult to complete the partition. From Eq. (6.109), it follows that

(6.118)

In analogy to Eqs. (6.111), (6.112), and (6.114), the following definitions are made:

(6.119)

from which it follows that

(6.120a)

and

(6.120b)

Here, Cff denotes the resistance coefficient associated with form drag and Hs denotes the extra depth (compared to the case of
skin friction alone) that results from form drag.

Up to this point, it is assumed the U, S, and ks are given. If, for example, H also is known, can be calculated from Eq. (6.110).
After Hs, Cfs, and τ bs are computed from Eqs. (6.113) to (6.115), it is possible to compute τbf, Hf, and Cff from Eqs. (6.116) and
(6.118).

6.8.2.2. Example of the Einstein partition.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Consider a sand-bed stream at a given cross section with a slope of 0.0004, a mean depth of 2.9 m, a value of median bed
sediment size of 0.35 mm, and a discharge per unit width of 4.4 m 2 /s. Assume that the flow is at near-normal conditions.
Compute values of τbs , τ bf, Cfs, Cff, Hs, and Hf.

Solution: U = 4.4/2.9 = 1.52 m/s. An appropriate estimate of ks for a sand-bed stream is

(6.121)

Solving Eq. (6.115) by successive approximation, it is found the Hs = 1.047 m. The following values then hold:

τbs = 4.11 newton’s/m2 = 0.725)

τbf = 7.27 newtons/m2 = 1.283)

τb = 11.38 newtons/m 2 = 2.008)

Cfs = 0.00178

Cff = 0.00315

Cf = 0.00493 = 14.5)

Hs = 1.047 m

Hf = 1.842 m

H = 2.9 m

In the above relations,

(6.122)

denotes a form Shield stress. In the above case, only some 30% of the total Shields stress (skin + form) contributes to moving
sediment.

The Einstein method provides a way of partitioning the boundary shear stress if the flow is known. It does not provide a direct
means of computing form drag. A method proposed by Nelson and Smith (1989) overcomes this difficulty.

6.8.2.3. Nelson-Smith partition.


Nelson and Smith considered flow over a dune; the flow is taken to separate in the lee of the dune. On the basis of experimental
observations, they use the following relation for form drag:

(6.123)

Here, Df denotes that portion of the streamwise drag force Dfs that is caused by form drag, B is the channel width, and Ur
denotes a reference velocity to be defined below. They evaluate the drag coefficient cD as

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.124)

It follows that

(6.125)

The reference velocity Ur is defined to be the mean velocity that would prevail between z = ks and z = Δ if the bedforms were not
there. From the logarithmic profile represented by Eq. (6.7), this is found to be given by

(6.126)

It is now assumed that a rough logarithmic law with roughness ks prevails from z = ks to z = D, and a different rough logarithmic
law with roughness kc prevails from z = D to z = H. Here kc represents a composite roughness length, including the effects of
both skin friction and form drag. The two laws are thus

(6.127a)

and

(6.127b)

Nelson and Smith (1989) matched the above two laws at the level z = Δ. After some manipulation, it is found that

(6.128)

The partition requires a prior knowledge of total boundary shear stress as well as roughness height ks, dune
height Δ, and dune wavelength Λ. Between Eqs. (6.123) and (6.124),

(6.129)

This equation can be solved for τbs , and thus τbf. The value of kc is then obtained from Eq. (6.128).

6.8.2.4. Example of the Nelson-Smith Partition.


The example is chosen to be rather similar to the previous one: H = 2.9 m, S = 0.0004, ks = 2.5 D50, D50 = 0.35 mm, Δ = 0.4 m, and
Λ = 15 m. The technique, which requires no iteration, yields the following results:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
τbs = 4.45 newtons/m2 = 0.785)

τbf = 6.93 newtons/m2 = 1.223)

= 11.38 newtons/m 2 (τ* = 2.008)

Kc = 0.0311 m

Cfs = 0.00130

Cff = 0.00203

Cf = 0.00333 = 17.3)

Hs = 1.134 m

Hf = 1.766 m

H = 2.9 m

In computing friction coefficients, the following relationship was used for the depth-averaged flow velocity:

(6.130)

The Nelson-Smith method does not require the assumption of quasi-normal flow.

6.8.3. Empirical Formulas for Stage-Discharge Relations


To use either the Einstein or Nelson-Smith partitions, it is necessary to know in advance the total effective boundary shear
stress In general, this is not known. As a result, the relations in themselves cannot be used to predict the boundary shear
stress (as well as the contributions from skin friction and form drag), and thus depth H, for a flow of, say, given slope S and
discharge per unit width qw.

A number of empirical techniques have been proposed to accomplish this. Only three are presented here; they are known to
perform well for sand-bed streams with dune resistance.

6.8.3.1. Einstein-Barbarossa Method.


The method of Einstein and Barbarossa (1952) is applicable for the case of dune resistance in a sand-bed stream. It assumes
an empirical relation of the following form:

(6.131)

Here,

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.132)

The Einstein-Barbarossa plot is shown in Fig. 6.23. Note that it implies that Cff declines for increasing That is, the relation
applies in the range for which increased intensity of flow causes a decrease in form drag.

Figure 6.23 Flow resistance due to bedforms. [after Einstein et al. (1952).]

In the Einstein-Barbarossa method, Cfs is computed from a relation similar to Eq. (6.113). That relation is used here to illustrate
the method, which uses the Einstein partition for skin friction and form drag.

6.8.3.2. Application of the Einstein-Barbarossa Method.


The Einstein-Barbarossa method is now used to synthesize a depth-discharge relation: that is, a relation betweenH and water
discharge Q is obtained. It is assumed that the river slope S and the sizes D50 and D35 are known. The river is taken to be wide
enough so that the hydraulic radius Rh H; otherwise, Rh should be used in place of H. In addition, the cross-sectional shape is
known, allowing for specification of the following geometric relation:

(6.133)

(It also is assumed that auxiliary relations for area A, wetted perimeter P and Rh as functions of H are known.)

A range of values of Hs is arbitrarily assumed, ranging from an extremely shallow depth to nearly bankfull depth (recall that Hs ≤
H). For each value of Hs, the calculation proceeds as follows:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
H → C fs Eq. (6.115)

Cfs, H s → U Eq. (6.116b)

Eq. (6.116b), (6.132)

Eq. (6.131); use the diagram

Cff U → H f Eq. (6.119)

H = Hs − H f Eq. (6.120b)

Q = UH · B(H) Eq. (6.133)

The result can be plotted in terms of H versus Q.

The analysis can be continued for bedload transport rates. That is, the parameter τbs can be computed from

(6.134)

and this parameter can be substituted into an appropriate bedload transport equation to obtainq. The volume bedload
transport rate Qb is then computed as

(6.135)

6.8.3.3. Engelund-Hansen Method.


The method of Engelund and Hansen (1967) also applies specifically for sand-bed streams. It is generally more accurate than
the method of Einstein and Barbarossa, to which it is closely allied. The method assumes quasi-uniform material; it is
necessary to know only a single grain size D. Roughness height ks is computed from Eq. (6.121).

The method uses the Einstein partition. Skin friction is computed usingEq. (6.112). Form drag is computed from the following
empirical relation:

(6.136)

where

(6.137a, b)

Equation (6.134) is shown graphically in Fig. 6.24. It has two branches, each corresponding to lower-regime and upper-regime
flows. The two do not meet smoothly, implying the possibility of a sudden transition. The point of transition is not specified,
which suggests the possibility of double-valued rating curves. The lower-regime branch of Eq. (6.136) is given by

(6.138)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The upper branch satisfies the relation

(6.139)

over a range; this implies an upper-regime plane bed. For higher values of Shields stress, τ* again exceeds implying antidune
resistance.

6.8.3.4. Application of the Engelund-Hansen Method.


The procedure parallels that of Einstein-Barbarossa relatively closely. It is assumed that the values of S and D as well as the
cross-sectional geometry are known. Values of Hs are selected, ranging from a low value to near bankfull. The calculation then
proceeds as follows:

H s → C fs → U Eq. (6.115) and (6.116b)

Eq. (6.116b), (6.137b)

Eq. (6.134); use Equation (6.136) or plot

Eq. (6.137b) and (6.116a)

Q = UH · B(H) Eq. (6.133)

The value of can then be used to calculate bedload transport rates in a fashion that is completely analogous to the procedure
outlined for the Einstein-Barbarossa method.

6.8.3.5. Brownlie method.


There are almost as many empirical resistance predictors for rivers as there are sediment transport relations. A fairly
comprehensive summary of the older methods can be found in ASCE Manual No. 54 (Vanoni, 1975). A recent empirical method
offered by Brownlie (1981a) has proved to be relatively accurate. It does not involve a decomposition of bed shear stress;
instead it gives a direct predictor of depth-discharge relations.

The complete method can be found in Brownlie (1981a), where the relation is presented for the case of lower-regime dune
resistance in a sand-bed stream. It takes the form

(6.140)

where σg denotes the geometric standard deviation of the bed material, and denotes a dimensionless water discharge per
unit width, given by

(6.141)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.24 Relation between grain shear stress and total shear stress (after Engelund and Hansen,
1976)

For known S, D50, and σg, qw, and thus Q = qw ·B is computed directly as a function of depth H.

6.9. SUSPENDED LOAD


6.9.1. Mass Conservation of Suspended Sediment
Suspended sediment differs from bedload sediment in that it can be diffused throughout the vertical column of fluid via
turbulence. Here, the local mean volume concentration of suspended sediment is denoted as . As long as the suspended
sediment under consideration is coarse enough not to undergo Brownian motion (i.e., silt or coarser), molecular effects can be
neglected. Suspended particles are transported solely by convective fluxes.

For an arbitrary volume of sediment-water mixture in the water column, the equation of mass balance of suspended sediment
can be written in words as

(6.142)

Insofar as the choice of volume V is entirely arbitrary, the following sediment conservation equation, averaged over turbulence-
induced fluctuations about the mean, can be obtained:

(6.143)

where and are the mean flow velocities in the s, n, and z directions, respectively, and the terms and are
sediment fluxes caused by turbulence, also known as Reynolds fluxes. The simplest closure assumption for these terms is

(6.144a)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.144b)

and

(6.144c)

where the kinematic eddy diffusivity Dd is assumed to be a scalar quantity. To solve Eq. (6.143), boundary conditions are
needed.

6.9.2. Boundary Conditions


Equation (6.143), when closed with a Fickian assumption, such as Eq. (6.144a, b, and c), represents an advection-diffusion
equation for suspended sediment. The condition of vanishing flux of suspended sediment across (normal to) the water surface
defines the upper boundary condition.

If uniform steady flow over a flat (when averaged over bedforms) bed is considered, the surface boundary condition for the net
vertical flux of sediment reduces to

(6.145)

where

(6.146)

is the net vertical flux of sediment.

The boundary condition at the bed differs from the one at the water surface because it must account for entrainment of
sediment into the flow from the bed and for deposition from the flow onto the bed. The mean flux of suspended sediment onto
the bed is given by –D, where

(6.147)

denotes the volume rate of deposition of suspended sediment per unit time per unit bed area. Here denotes a near-bed value
of

The component of the Reynolds flux of suspended sediment near the bed that is directed upward normal to the bed may be
termed the rate of erosion, or more accurately, entrainment of bed sediment into suspension per unit bed area per unit time. The
entrainment rate E is thus given by

(6.148)

where w′ and c′ denote turbulent fluctuations around both the mean vertical fluid velocity and the mean sediment concentration,
respectively. The “overbar” denotes averaging over turbulence. The term near bed used to avoid possible singular behavior at
the bed (located at z = 0).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
It is seen from the above equations that the net upward normal flux of suspended sediment at (or rather just above) the bed is
given by

(6.149a)

where

(6.149b)

denotes a dimensionless rate of entrainment of bed sediment into suspension. The required bed boundary condition, then, is a
specification of Es. Typically, a relation of the following form is assumed:

(6.150a)

where τbs denotes the boundary shear stress caused by skin friction.

Furthermore, it is assumed that an equilibrium steady, uniform suspension has been achieved. It follows that there should be
neither net deposition on nor erosion from the bed. That is, = 0, yielding

(6.150b)

This relation simply states that the entrainment rate equals the deposition rate; thus, there is no net normal flux of suspended
sediment at the bed.

6.9.3. Equilibrium Suspension in a Wide Rectangular Channel


Consider normal flow in a wide, rectangular open channel. The bed is assumed to be erodible and has no curvature when
averaged over bedforms. The z-coordinate is quasi-vertical, implying low channel slope S. Similarly, the suspension is assumed
to be in equilibrium. That is, is a function of z alone (Fig. 6.25). The flow and suspension are uniform in s and n and steady in
time; thus, Eq. (6.143) reduces to

(6.151)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.25 Definition diagram for sediment entrainment and deposition

It is appropriate to close this equation with the assumption of an eddy diffusivity, as inEq. (6.144c); thus, Eq. (6.151) becomes

(6.152)

Equation (6.152) has a simple physical interpretation. The term –vs represents the rate of sedimentation of suspended
sediment under the influence of gravity; it is always directed downward. If all the sediment is not to settle out, there must be an
upward flux that balances this term. The upward flux is provided by the effect of turbulence, acting to yield a Reynolds flux.
According to Eq. (6.144c), this flux will be directed upward as long as d /dz < 0. It follows that the equilibrium suspended-
sediment concentration decreases for increasing z: therefore, turbulence diffuses sediment from zones of high concentration
(near the bed) to zones of low concentration (near the water surface).

6.9.4. Eddy Diffusivity


Further progress requires an assumption for the kinematic eddy diffusivityDd . The simple approach taken here is that of Rouse
(1957), which involves the use of the Prandtl analogy. The argument is as follows: Fluid mass, heat, momentum, and so on
should all diffuse at the same kinematic rate because of turbulence and thus have the same kinematic eddy diffusivity because
each is a property of the fluid particles, and the fluid particles are what is being transported by Reynolds fluxes.

Although the Prandtl analogy is by no means exact, it has proved to be a reasonable approximation for many turbulent flows. Its
application to sediment is more of a problem. Inertial effects might cause the sediment particles to lag behind the fluid,
resulting in a lower eddy diffusivity for sediment than for the fluid. Furthermore, the mean fall velocity of sediment grains should
reduce their time of residence in any given eddy, again reducing the diffusive effect. If the particles are not too large, however, it
may be possible to equate the vertical diffusivity of the sediment with the vertical eddy viscosity (eddy diffusivity of momentum)
of the fluid as the first approximation. This is done here.

The velocity profile is approximated as logarithmic throughout the depth. To account for the possible existence of bedforms,
the turbulent rough law embodied in Eq. (6.127b) is used:

(6.153)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Here kc is a composite roughness chosen to include the effect of bedforms, as outlined inSec. 6.8.2.3. Furthermore, according
to Eq. (6.2), the bed shear stress is given by

(6.154)

where b is chosen to be close to the bed: i.e.,

(6.155)

Now the kinematic eddy viscosity Dd is defined as

(6.156)

where the distribution of fluid shear stresses τ is given by

(6.157)

From the above equations, it is quickly found that

(6.158)

where κ = 0.4 is Von Karman’s constant.

The above relation is the Rousean relation for the vertical kinematic eddy viscosity. The form predicted is parabolic in shape.
Although strictly applying to the turbulent diffusion of fluid momentum, it is equated to the eddy diffusivity of suspended
sediment mass below. If Dd is averaged in the vertical, the following result is obtained:

(6.159)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.26 Vertical suspended sediment distribution (after Vanoni, 1961).

This relation is useful to estimate the longitudinal dispersion of fine-grained sediment in rivers and streams.

6.9.5. Rousean Distribution of Suspended Sediment


The nominal “near bed” elevation in applying the bottom boundary condition is taken to bez = b, where b is a distance taken to
be extremely close to the bed: i.e., satisfying condition Eq. (6.155). In the Rousean analysis, this value cannot be taken as z = 0
because Eq. (6.153) is singular there.

Equation (6.158) is now substituted into Eq. (6.152), which is then integrated from the nominal bed level to distancez above the
bed in z. The resulting form can be cast as

(6.160)

where Z denotes the Rouse number, a dimensionless number given by

(6.161)

Further reduction yields the following profile:

(6.162)

Some sample profiles of suspended sediment plotted in Rousean form are provided inFig. 6.26.

Note that from Eq. (6.150b), is equal to the dimensionless sediment entrainment rate Es in the case of the present
equilibrium suspension. This provides an empirical means to evaluate Es as a function of τbs and other parameters, as will be

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
s bs
shown.

6.9.6. Vertically Averaged Concentrations: Suspended Load


Assuming that a value of near-bed elevation b is chosen approximately, Eq. (6.162) can be used to evaluate a depth-averaged
volume suspended-sediment concentration C, defined by

(6.163)

Using Eq. (6.162), then

(6.164a)

where

(6.164b, c)

In the above relation, the integral is evaluated easily by means of numerical techniques. Einstein (1950) representedI1
in the form

(6.165)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.27a Function I 1 in terms of ξb = b/H for values of Z:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.27b Function –I2 in terms of ξb = b/H for values of Z:

where I1 is given in tabular form in the attached Fig. 6.27a.

The streamwise suspended load qs was seen in Eq. (6.61a) to be given by the relation

(6.166)

Reducing with the aid of Eqs. (6.153) and (6.162), we find that

(6.167)

Here,

(6.168)

The integral I2 is again evaluated easily numerically: Einstein provides the relation

(6.169)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where –I 2 is given in tabular form in Fig. 6.27b. Brooks (1963) also proposed an interesting way to calculate suspended load
discharge from velocity and concentration parameters.

It is apparent that further progress is predicated on a method for evaluating the “reference concentration” , or equivalently
(for the case of equilibrium suspensions) the sediment entrainment rate Es Such a relation is necessary to model transport of
suspended sediment (e.g., Celik and Rodi, 1988).

6.9.7. Relation for Sediment Entrainment


A number of relations are available in the literature for estimating the entrainment rate of sediment into suspensionEs (and thus
the reference concentration for the case of equilibrium). Table 6.5 summarizes all the relations that are available. García and
Parker (1991) performed a detailed comparison of eight such relations against data. The relations were checked against a
carefully selected set of data pertaining to equilibrium suspensions of uniform sand. In this case, it is possible to measure
directly at some near-bed elevation z = b, and to equate the result to Es according to Eq. (6.150b)

The data consisted of some 64 sets from 10 different sources, all pertaining to laboratory suspensions of uniform sand with a
submerged specific gravity R near 1.65. Information about the bedforms was typically not sufficient to allow for a partition of
boundary shear stress in accordance with Nelson and Smith (1989). As a result, the shear stress caused by skin friction alone
τbs and the associated shear velocity caused by skin friction u*s, given by

(6.170)

were computed using Eq. (6.114) and the following relation for ks,

(6.171)

or a similar method.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Table 6.5 Existing formulas to estimate sediment entrainment or near-bed concentration under equilibrium conditions.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.28 Sediment entrainment function (after García and Parker, 1991).

The data covered the following ranges:

Es: 0.0002 ~ 0.06

u*s /vs: 0.70 ~ 7.50

H/D: 240 ~ 2400

Rep 3.50 ~ 37.00

The range of values of Re p corresponds to a grain size ranging from 0.09 mm to 0.44 mm. Except for the relatively small values
of H/D, the values cover a range that includes typical field sand-bed streams.

Three of the relations for Es performed particularly well and are presented here. The first is the relation of García and Parker
(1991). The reference level is taken to be 5 percent of the depth: that is,

(6.172)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.29 Comparison of predicted and observed near-bed concentration for García-Parker
function

The good performance of this relation is not overly surprising because the relation was fitted to the data. The relation takes the
form

(6.173a)

where

(6.173b)

and

(6.173c)

Equation (6.173a) is compared against the data in Fig. 6.28. Predicted values of Es are compared with observed values in Fig.
6.29.

A second relation that performed well is that of Van Rijn (1984), which takes the form

(6.174)

where denotes the Shields stress caused by skin friction, given by Eq. (6.112). For the purposes of the present comparison, b
was again set equal to 5 percent of the depth: i.e., Eq. (6.172) was used. Van Rijn computed τbs from relations that are similar
to Eqs. (6.115) and (6.116b). Van Rijn's relations are

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.175a)

where, for uniform material,

(6.175b)

Note that in Eq. (6.175), the total depth H is used, in contrast to Eq. (6.115) where Hs is used.

In performing the comparison, García and Parker (1991) estimated from a fit to the Shields curve due to Brownlie (1981a).
This fit is given by Eq. (6.44). Predicted and observed values of Es are presented in Fig. 6.30.

Figure 6.30 Comparison of predicted and observed near-bed concentration for van Rijn function

Figure 6.31 Comparison of predicted and observed near-bed concentration for Smith-McLean
function.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
A third relation that performs well is that of Smith and McLean (1977) which can be expressed as

(6.176a)

where

(6.176b)

The value b at which E is to be evaluated is given by the following relation:

(6.176c)

where ks denotes the equivalent roughness height for a fixed bed.

For the purpose of comparison, García and Parker used Eq. (6.171) to evaluate ks and used Eq. (6.115) to evaluate τbs . Critical
Shields stress was evaluated with Eq. (6.44). Predicted and observed values of Es are shown in Fig. 6.31.

6.9.8. Entrainment Relation for Sediment Mixtures


García and Parker (1991) provided a generalized treatment for the entrainment rate in the case of mixtures. Let the grain-size
range of bed material be divided into N subranges, each with mean size on the phi scale and geometric mean diameter
where j = 1...N. Let Fj denote the volume fraction of material in the surface layer of the bed in thejth grain range. In
analogy to Eq. (6.148), it is assumed that

(6.177a)

where Ej denotes the volume entrainment rate for the jth subrange and the functional relation between Es and Z uj is given by Eq.
(6.173a). The parameter Z uj is specified as

(6.177b)

In the above relations, vsj denotes the fall velocity of grain size Dj in quiescent water, D50 denotes the median size of the surface
material in the bed

(6.177c)

and the parameter λm is given by

(6.177d)

Here, denotes the arithmetic standard deviation of the bed surface material on the phi scale, given byEq. (6.30).

The García-Parker relation for mixtures reduces smoothly to the relation for uniform material in the limit as . It was
developed and tested with three sets of data from two rivers: the Rio Grande and the Niobrara River. Recently, the García-Parker

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
formulation also has been used to interpret observations of sediment entrainment into suspension by bottom density currents
(García and Parker, 1993).

6.9.9. Example of Computation of Sediment Load and Rating Curve.


Consider the example of a stream in Sec. 6.8.2. For this stream, S = 0.0004 and D = 0.35 mm (uniform material). At bankfull
flow, the stream width is 75 m. For flows below bankfull, the following relation holds:

where the subscript bf denotes bankfull. Assume that the stream is wide enough to equate the hydraulic radius Rh with the
cross-sectionally averaged depth H.

Compute the depth-discharge relations for flows up to bankfull (lower regime only) using the Engelund-Hansen method. PlotH
versus Q. Use the results of the Engelund-Hansen method to compute values of as well.

Use the values of to compute the bedload discharge Qb = qb · B using the Ashida-Michiue formulation. For each value of H
and U, back-calculate the composite roughness kc. Then compute the suspended load Qs = qs · B from the Einstein formulation
and the relation for Es by García and Parker. Plot Qb, Qs, and QT = Qb + Qs as functions of water discharge Q.

Solution: In this example, the flow depth, bedload discharge, and suspended load discharge are computed as a function of flow
discharge for a stream with the following properties:

S= 0.0004

Ds = 0.35 mm = 3.5 × 10-4 m

R= 1.65

B= 75 m at bankfull

H= 2.9 m at bankfull

The calculations are performed for flows up to bankfull. For flows below bankfull, the following relation is used to calculate the
stream width:

(i)

where the subscript bf indicates bankfull values. Solving for the stream width B,

(ii)

The methods used to determine Q, Qb, Qs, and Qbf are described below. A computer program can be written, or a spreadsheet
can be used, to perform the necessary calculations. All computations and results are summarized in Table 6.6.

6.9.9.1. Depth-discharge calculations.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The depth-discharge relation is computed using the Engelund-Hansen method. The calculations are performed by assuming a
value for Hs (the flow depth that would be expected in the absence of bedforms), then calculating the actual flow depth (H) and
the flow discharge (Q). Hs is varied between 0.22 m and the bankfull value of 2.9 m. The first step in calculating the depth-
discharge relation is to compute the resistance coefficient caused by skin drag (Cfs) from Hs:

(iii)

Table 6.6 Computation of Total Sediment Load.

where κ is the von Karman constant (0.4) and ks is given by

(iv)

The depth-averaged flow velocity (U) can be found from Cfs and Hs:

(v)

The Shields stress caused by skin friction is given by

(vi)

According to Engelund-Hansen, the total Shields stress for the lower regime can be found from the following relation:

(vii)

The flow depth can be calculated from the Shields stress as follows:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(viii)

Finally, the discharge can be calculated from the results of Eqs. (v) and (viii):

(ix)

where B must be adjusted according to Eq. (ii) for flows less than bankfull. A plot of the depth-discharge relation is shown in
Fig. 6.32.

6.9.9.2. Bedload discharge calculations.


The dimensionless bedload transport rate (q*) is found from the Ashida-Michiue formulation:

(x)

where is calculated in Eq. (vi) and is taken to be 0.05. The bedload transport rate per unit width (qb) is given by

(xi)

Therefore, the bedload transport rate (in m3 /s) is given by

(xii)

Again, B must be adjusted according to Eq. (ii) for flows less than bankfull.

6.9.9.3. Sediment load discharge calculations.


The Einstein formulation is used to compute the suspended load transport rate per unit width (qs):

(xiii)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.32 Example of flow discharge rating curve

where

(xiv)

If the suspension is assumed to be at equilibrium, = Es. The dimensionless rate of entrainment (Es) is calculated with the
relation of García and Parker (1991):

(xv)

where A is equal to 1.3 × 10−7 and

(xvi)

(xvii)

and

(xviii)

Notice that for the entrainment formulation, the shear velocity associated with skin friction u*s must be used. The temperature
is assumed to be about 20°C; therefore, the kinematic viscosity is about 10–6 m2 /s. An iterative method, or Eq. (6.38), is used to
calculate the terminal fall velocity of the sediment particles vs, which is found to be 5.596 × 10 –2 m/s. The composite
roughness (kc) is calculated according to the following relation:

(xix)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The parameters I1 and I2 are found by numerical integration of the following equations:

(xx)

and

(xxi)

where is taken to be 0.05 and

(xxii)

The numerical integrations can be performed with Numerical Recipes subroutines (Press et al., 1986) or can be obtained from
Figs. 6.27a and b. The suspended load transport rate per unit width calculated according to Eq. (xiii) is used to compute the
suspended load transport rate (in m3 /s):

(xxiii)

For flows less than bankfull, B must be adjusted according to Eq. (ii).

6.9.9.4. Determination of bankfull flow discharge (Qbf).


The flow discharge at bankfull (Qbf) is determined by assuming that up to bankfull flow, lower regime conditions exist. The
bankfull flow depth for this stream is assumed to be 2.9 m. Then, for bankfull flow, the total shear stress τ* is

(xxiv)

From Engelund-Hansen,

(xxv)

(xxvi)

(xxvii)

(xxviii)

and

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(xxix)

A plot of Qb, Qs, and QT = Qb + Qs as functions of water discharge are shown in Fig. 6.33. For flows up to 100 m3 /s, the bedload
discharge is larger than the suspended load discharge. As the flow discharge increases, the suspended load is much larger than
the bedload all the way up to bankfull flow conditions.

Also notice that the composite roughness kc increases first with flow discharge for low flows but, from then on, decreases
monotonically as the bedforms begin to be washed out by the flow. For bankfull conditions, the bedforms have a small effect on
flow resistance in this particular problem.

6.10. DIMENSIONLESS RELATIONS FOR TOTAL BED


MATERIAL LOAD IN SAND-BED STREAMS
6.10.1. Form of the Relations
In the analysis presented in previous sections, the guiding principle has been the development of mechanistically accurate
models of the bedload and suspended load components of bed material load. The total bed material load is then computed as
the sum of the two. That is, where q denotes the volume bedload transport rate per unit width and qS denotes the volume
suspended load transport rate per unit width (bed material only), the total volume transport rate of bed material per unit width is
given by

(6.178)

Figure 6.33 Example sediment discharge rating curves for bedload, suspended load, and total load

Another simpler approach is to ignore the details of the physics of the problem and instead use empirical techniques, such as
regression analysis, to correlate dimensionless parameters involving qt to dimensionless flow parameters inferred to be
important for sediment transport. This can be implemented in the strict sense only for equilibrium or quasi-equilibrium flows:
i.e., for near-normal conditions. The resulting relations are no better than the choice of dimensionless parameters to be
correlated. They also are less versatile than physically based relations because their application to nonsteady, nonuniform flow
fields is not obvious. On the other hand, they have the advantage of being relatively simple to use and of having been calibrated
to sets of both laboratory and field data often deemed to be trustworthy.

Here, four such relations are presented, those of Engelund and Hansen (1967), Brownlie (1981a), Yang (1973), and Ackers and

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
White (1973). They apply only to sandbed streams with relatively uniform bed sediment. The first two relations are the most
complete because each is presented as a pair of relations for total load and hydraulic resistance. The latter two are presented
as relations for total load only. In most cases, it will be necessary for the user to specify a relation for hydraulic resistance as
well to perform actual calculations; the latter relations for load give no guidelines for this.

The importance of using transport and hydraulic resistance relations as pairs cannot be overemphasized. Consider, for
example, the simplest generalization beyond the assumption of normal flow: i.e., the case of quasi-steady, gradually varied, one-
dimensional flow. The governing equations for a wide rectangular channel can be written as

(6.179a)

(6.179b)

where the friction slope S f is given as

(6.180)

A slightly more general form for nonrectangular channels is

(6.181a)

(6.181b)

where A is the channel cross-sectional area and the friction slope S f is given as

(6.182)

In the above equations, Rh denotes the hydraulic radius and denotes the water surface elevation above the deepest point in
the channel.

Note that in the case of normal flow, the momentum equations reduce toS f = S, or τ b = ρgHS for the wide rectangular case and
τb = ρgRhS for the nonrectangular case. However, in the case of gradually varied flow,S f ≠ S; in this case, the bed slope S
cannot be used as a basis for calculating sediment transport. The appropriate choice is S f, so that from Eq. (6.182), for
example,

(6.183)

For the case of gradually varying flow, then, it should be apparent that the friction slope necessary to perform sediment
transport calculations must be obtained from a predictor of hydraulic resistance.

A few parameters are introduced here. Let Q denote the total water discharge and Qst denote the total volume bed material
sediment discharge. Furthermore, let Ba denote the “active” width of the river over which bed material is free to move. In
general, Ba is usually less than water-surface width B as a result of the common tendency for the banks to be cohesive,
vegetated, or both. Thus, it follows that

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.184a)

and

(6.184b)

One dimensionless form for dimensionless total bed material transport is

(6.185)

where D is a grain size usually equated to D50. Another commonly used measure is concentration by weight in parts per million,
here called Cs, which can be given as

(6.186)

6.10.2. Engelund-Hansen Relations


6.10.2.1. Sediment transport.
This relation is among the simplest to use for sediment transport and also among the most accurate. It was determined for a
relatively small set of laboratory data, but it also performs well as a field predictor. It takes the form

(6.187)

where Cf is the total resistance coefficient (skin friction plus form drag) and τ* denotes the total (skin friction plus form drag)
Shields stress based on the size D50.

6.10.2.2. Hydraulic resistance.


The hydraulic resistance relation of Engelund and Hansen (1967) has already been introduced; it must be written in several
parts. The key relation for skin friction is

(6.188a)

where ks = (2 ~ 2.5) · D50. Here, Rhs denotes the hydraulic radius caused by skin friction, which often can be approximated byHs.
The relation for form drag can be written in the following form:

(6.188b)

where for lower regime,

(6.188c)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
and for upper regime,

(6.188d)

An approximate condition for the transition between lower and upper regime is

(6.188e)

Computational procedure for normal flow. The water discharge Q, slope S, and grain size D50 must be known. In addition,
channel geometry must be known so that B, Ba, A, H, P, and Rh are all known functions of stage (water-surface elevation) The
procedure is best outlined assuming that Rhs is known and that Q is to be calculated, rather than vice verse. For any given value
of Rhs (or Hs), U can be computed from Eq. (6.188a). Noting that and τ* = RhS/(RD50), τ*, and thus Rh can be
computed from Eq. (6.188b-e). The plot of Rh versus is used to determine which is then used to determine B, Ba, H, A, P,
and so on. Discharge Q is then given by Q = UBH. In actual implementation, this process is reversed (Q is given and Rhs and so
forth are computed). This requires an iterative technique; Newton-Raphson is not difficult to implement.

Once the calculation of hydraulic resistance is complete, it is possible to proceed to the computation of total bed material load
Qst. The friction coefficient Cf is given by (gRhS)/U2 . Putting the known values of Cf and τ* into (6.187), qt*, and thus qt can be
computed. It follows that Qst = qtBa.

Computational procedure for gradually varied flow. To implement the method for gradually varied flow, it is necessary to recast
the above formulation into an algorithm for friction slope S f, which replaces S everywhere in the formulation of Eqs. (6.188a-e).
The formulation is then solved in conjunction with Eqs. (6.179a and b) or Eqs. (6.181a and b) to determine the appropriate
backwater curve. Once Cf and τb are known everywhere, the sediment transport rate can be calculated from Eq. (6.187).

6.10.3. Brownlie Relations


6.10.3.1. Sediment transport.
The Brownlie relations are based on regressions of more than 1000 data points pertaining to experimental and field data. For
normal or quasi-normal flow, the transport relation takes the form

(6.189a)

where

(6.189b)

(6.189c)

(6.189d)

and

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.189e)

In Eq. (6.189a), cf = 1 for laboratory flumes and 1.268 for field channels. The parameters and Re p are the ones previously
introduced in this chapter.

6.10.3.2. Hydraulic resistance.


The Brownlie relations for hydraulic resistance were determined by regression from the same set of data used to determine the
relation for sediment transport. The relation for lower regime flow is

(6.190a)

The corresponding relation for upper regime flow is

(6.190b)

In the above relations,

(6.190c)

The distinction between lower and upper regime is made as follows. For S > 0.006, the flow is always assumed to be in upper
regime. For S < 0.006, the largest value of Fg at which lower regime can be maintained is taken to be

(6.190d)

and the smallest value of Fg for which upper regime can be maintained is taken to be

(6.190e)

In the above relations,

(6.190f)

6.10.3.3. Computational procedure for normal flow.


It is necessary to know Q, S, D50, σg, and cross-sectional geometry as a function of stage. The computation is explicit, although
trial and error may be required to determine the flow regime. Hydraulic radius is computed from Eq. (6.190a) or Eq. (6.190b),
and the result can be substituted into Eq. (6.189a) to determine the concentration Cs in parts per million by weight. The
transport rate Qst is then computed from Eq. (6.186).

6.10.3.4. Computational procedure for gradually varied flow


The Brownlie relation is not presented in a form which obviously allows for extension to gradually varied flow. The most

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
unambiguous procedure, however, is to replace S with S f in the resistance relation, and couple it with a backwater calculation
on order to determine Sf . The friction slope is then substituted into Eq. (6.189a) in place of the bed slope in order to determine
the sediment transport rate.

6.10.4. The Ackers–White Relation


Like the Brownlie relation for sediment transport, this relation is based on a massive regression. Several years after is was
presented, a corresponding relation for hydraulic resistance was also presented. The relation for hydraulic resistance, however,
does not appear to be among the best predictors. As a result, only the load equation is presented here. It takes the form

(6.191a)

where

(6.191b,c)

The parameters n, and Aaw are determined as a functions of Dgr, where

(6.191d)

in the following fashion. If Dgr > 60, then

(6.191e and f)

and

(6.191g and h)

If 1 < Dgr < 60, then

(6.191i and j)

(6.191k)

(6.191)

Note that all logarithms here are base 10, and u* retains its previously introduced meaning as shear velocity.

6.10.5. Yang Relation

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
This relation also was determined by regression. Its form is

(6.192a)

where

(6.192b)

(6.192c)

and Uc denotes a critical flow velocity given by

(6.192d)

Not that the logarithms are all base 10 and that vS retains its previous meaning as fall velocity.

6.10.6. Comparison of the Relations Against Data


In the following eight diagrams (Fig. 6.34a-h) taken from Brownlie (1981b) all four relations are compared against first
laboratory, then field data. The plots are in terms of the ratio of calculated versus observed concentration as a function of
observed concentration Cs in parts per million by weight. In the case of a perfect fit, all the data would fall on the line
corresponding to a ratio of unity. The middle dotted line on each diagram shows the median value of this ratio; the upper and
lower dotted lines correspond to the 84th percentile and the 16th percentile. The closer the median value is to unity and the
smaller the spread is between the two dotted lines, the better is the predictor.

The Engelund-Hansen relation is seen to be a good predictor of both laboratory and field data despite its simplicity. The
Brownlie relation gives the best fit of both the laboratory and field data shown. This is partly to be expected because the relation
was determined by regressing against the data shown in the figures. The Ackers-White relation predicts the laboratory data
essentially as well as the Brownlie relation does, but its predictions of field data are relatively low. The Yang equation does a
good job with the laboratory data but a rather poor job with the field data.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.34a Ratio of concentration calculated by the Engelund and Hansen (1967) technique to
observed concentration as a function of observed concentration, for field data.

Figure 6.34b Ratio of concentration calculated by the Engelund and Hansen (1967) technique to
observed concentration as a function of observed concentration, for laboratory data.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.34c Ratio of concentration calculated by the Brownlie (1981b) technique to observed
concentration as a function of observed concentration, for laboratory data.

Figure 6.34d Ratio of concentration calculated from Brownlie (1981b) technique to observed
concentration as a function of observed concentration, for field data.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.34e Ratio of concentration calculated by the Ackers and White (1973) technique to
observed concentration as a function of observed concentration, for laboratory data.

Figure 6.34f Ratio of concentration calculated by the Ackers and White (1973) technique to
observed concentration as a function of observed concentration, for field data.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.34g Ratio of concentration calculated by the Yang (1973) technique to observed
concentration as a function of observed concentration, for laboratory data.

Figure 6.34h Ratio of concentration calculated by the Yang (1973) technique to observed
concentration as a function of observed concentration, for field data.

6.11. HYDRAULICS OF RESERVOIR SEDIMENTATION


6.11.1. Introduction
The construction of reservoirs allows for the controlled storage of water.To develop a successful reservoir, the characteristics
of the sites sediment transport must be considered. As a matter of course, water backed up behind a dam will experience a
marked decrease in sediment-carrying capacity. As a result, if site characteristics are correct, large quantities of sediment will
be deposited within the reservoir basin. Over time, the reservoir will, in effect, fill with sediment, greatly decreasing its storage
capacity. In 1988, Morris and Fan published an excellent handbook on reservoir sedimentation.

When designing a reservoir, it is important to predict the progress of sedimentation. In practice, these predictions are often
carried out using empirical and semiempirical methods that have been developed through observation and measurements of
operating reservoirs. Although these methods do provide helpful design information, the drawback is that they are not firmly

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
rooted in the physics of sediment transport. Instead, they provide a prediction based on a synthesis of past observations. As a
result, an engineer conducting these calculations easily loses touch with the basic mechanisms governing reservoir
sedimentation. Sedimentation is treated as a bulk process, and the relative role of bed-load versus suspended-load transport is
not always fully understood.

The following exercise presents a view of reservoir sedimentation based on theoretical relations. A gorgelike reservoir is
considered to allow for a 1-D model (Hotchkiss and Parker, 1991). The following conditions are given: the flow per unit width qw
= 1.427 m 2 /s is taken as constant. The stream has an initial slope S = 0.0003. The sediments mean diameter and fall velocity
are Ds = 0.3 mm and vs = 4.25 cm/s respectively. Suppose a reservoir is placed at some point on the river so that the water
surface is raised and held at an elevation equal to 10 m above the elevation of the initial bed at the dam site. Obviously, the dam
will generate a backwater effect, which in turn will reduce the flows ability to transport sediment through the reservoir. The
quasi-steady state approximation will be used to develop a model of reservoir sedimentation based on the governing equations
of conservation of momentum, bedload and suspended-load relations, and the Exner equation. The model will be used to
predict the level of reservoir sedimentation and delta progression for time intervals of 2, 5, 10, 20, and 30 years. First, the model
will be run considering bed-load transport only; second, suspended load also will be included to help identify the relative roles
these two forms of transport play.

The flow discharge per unit width qw used herein is equivalent to the “dominant” water discharge which, if continued constant
for an entire year, would yield the mean annual sediment discharge.

Of course, it is impractical to assume that a model as simple as the one presented here could replace the empirical methods of
predicting reservoir sedimentation. After all, a steady flow, 1-D, constant reservoir-elevation model seriously limits the model’s
application, and transport relations are not easily transposed from site to site. Still, the following provides an understanding of
the physical mechanisms causing reservoir sedimentation. An ideal reservoir-sedimentation model would be based in sediment
transport physics while respecting (and matching) the vast quantity of empirical observations available.

6.11.2. Theoretical Considerations


As in any sediment transport study, it is first necessary to identify the appropriate resistance and bedload transport relations
that hold for the site under consideration. For this model, the following relations have been chosen:

(6.193)

and

(6.194)

where τ* is the Shields Stress; is the critical Shields stress, which is taken to have a value of 0.03; h stands for the flow
depth; and ks is the roughness height. Cf is the resistance coefficient, and is the Einstein dimensionless bedload transport
defined below:

(6.195)

where qb is the volumetric bedload transport per unit width having the dimensions of m2 /s, where qb is the volumetric bedload
transport per unit width having the dimensions of m2 /s, R is the submerged specific gravity (taken as 1.65 for quartz), and D is
the mean diameter of the sediment particles.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
The following conservation relation can be used for suspended-load sediment routing:

(6.196)

where x is the coordinate in the streamwise direction, qw = Uh, C is the average volumetric suspended sediment concentration,
and ro C = = near-bed sediment concentration. The shape factor ro is given by the approximate relationship (Parker et al.,
1987):

(6.197)

Therefore, the suspended load transport (volume per unit width per unit time) through a section can be evaluated as the product
of the average sediment concentration and the flow discharge per unit width:

(6.198)

All that remains is to evaluate Es, the sediment-entrainment coefficient. This is accomplished with the García-Parker relation:

(6.199a)

where

(6.199b)

and

(6.199c)

With the above equations and the assumption of a rectangular cross section (qw = Uh), one can calculate the normal flow and
equilibrium transport conditions for the river. These calculations, shown next, will serve as the initial conditions for the
sedimentation study.

6.11.3. Computation of Normal Flow Conditions


From the Manning-Strickler relation (Eq. 6.194),

where ks = 2.5 Ds and

Now, qb can be computed:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Then

Estimation of C, Es and ro :

Then

For equilibrium conditions, entrainment and deposition rates are the same thus, with the help ofEq. (6.196),

and, finally, qs can be computed as

6.11.4. Governing Equations

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Next, it is necessary to identify the governing equations. Raising the water surface through a control structure results in the
development of backwater effects. The backwater profile can be calculated using the standard 1-D St. Venant equation
expressed in terms above parameters, with U being the flow velocity in x, the streamwise direction. The symbol η stands for the
elevation of the bed above the datum:

(6.200)

The backwater change in the water depth will cause a change in the transport of sediment. This phenomenon can be captured
using the Exner equation with λp being the porosity of the bed sediment (taken at 0.3):

(6.201)

Notice that a time differential appears in both of the above equations; this reflects the fact that both hydraulic and transport
conditions change continuously in time. The two equations are coupled through η, the bed elevation. Of course, a simultaneous
solution of both equations, including the time derivative, is difficult.

To simplify the model and expedite a solution, the quasi-steady-state approximation can be used. Not surprisingly, analysis has
shown, that the time scale for sedimentological changes is much larger than that for changes in flow condition. Simply put, if
the time changes of hydraulic conditions are driven by changes in sediment transport, they will occur slowly. Within an
appropriate time step, the flow conditions can be considered to be steady. In this way, it is possible to drop the time differential
in the St. Venant equation:

(6.202)

Equations (6.201) and (6.202), in conjunction with continuity (qw = Uh), provide the theoretical basis for the following analysis.
In the quasi-steady-state analysis, the backwater curve resulting from a forced raise in water elevation is calculated first, (Eq.
6.202). The new water depths for the time step are used to calculate a new bed position E ( q. 6.201), and these values are used
in the next time step to determine a new backwater profile. The procedure repeats for each time step.

6.11.5. Discussion of Method


As discussed in a previous section, initial “normal” flow conditions can be calculated through consideration of the resistance
and transport relations. Far away from the dam, where backwater effects are negligible, normal flow and equilibrium transport
conditions will exist.

A numeric scheme and computational grid must be chosen to evaluate the quasi-steady-state governing equations as they
relate to reservoir sedimentation. First, it is necessary to develop a spatial computational grid. The grid used in this numerical
experiment begins 40 km upstream of the front near the dam face. The length is divided into reaches of 200 m, resulting in 201
nodes to be evaluated. This length allows for initial backwater computation to very nearly reach the normal depth at the
upstream end.

Using this grid, it is possible to develop a numerical scheme for solving the governing equations. To begin the simulation, a
backwater calculation starting at the downstream end of the grid must be conducted. Combining the momentum equation (Eq.
6.202) and water continuity (qw = Uh) yields:

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.203)

where

(6.204)

and the friction slope S f is given by

(6.205)

If the value of h (and thus H) is known at node i + 1, its value at node i (upstream) can be calculated using the following finite
difference scheme:

(6.206)

The above expression can be expanded and written as a function of hi:

(6.207)

Now, a Newton-Raphson method can be used to evaluate hi. In this method, an arbitrary guess at hi can be refined by Δhi using
the expression

(6.208)

where

(6.209)

The computation begins at the downstream end of the problem, where the initial bed elevation must be specified according to
the normal flow conditions that existed before raising the water surface by 10 m. The first jump is Δx = 200 m, and each
subsequent jump is from node, to node with Δx = 200 m. Once the computation has progressed to the final node, the hydraulic
conditions for the initial bed condition are known. Variables such as U, qb , and qs can be calculated easily for each node once
the water depth is known.

Knowing the hydraulic conditions, the next necessary step is to evaluate the corresponding change in bed elevation. This is
accomplished with the Exner equation written in the backward finite difference form

(6.210)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where j is the current time step, j + 1 is the next time step, and i – 1 is the node immediately upstream of the node being
calculated. The calculation proceeds in the downstream direction. For the first node, the same technique is used, and the initial
“normal” bedload and suspended-load transport “feed rate” are used for the upstream values. It is crucial to choose a time step
that upholds the assumptions inherent to the quasi-steady-state approximation. Here, for bedload transport only, a time step of
0.01 year (3.65 days) is used, and for runs with both suspended and bedload transport, a time step of 0.001 year (365 days) is
used. These time steps are small enough to maintain theoretical integrity and numeric stability. The calculated elevations are
fed into the next time step for the adjustment of hydraulic conditions. The above procedure is continued for each time step.

To complete the numerous computations necessary for this procedure, a computer program must be written to facilitate the
numerical computations.

6.11.6. Results
The initial conditions for the water surface profile and bed elevation are shown inFig. 6.35. Changes in river profile with time
under the consideration of bedload transport only are shown in Fig. 6.36. The initial condition and the conditions after 2, 5, 10,
20, and 30 years are plotted. Figure 6.37 presents the same data, taking into account both bed and suspended sediment
transport. Not surprisingly, the delta formation is accelerated considerably when total load (bed and suspended) is considered.

Figure 6.35 Water surface elevation before and after reservoir construction

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.36 Development of delta for bedload only.

Figure 6.37 Development of delta for both bedload and suspended load

Both the heights and lengths of deltas are greater for total load calculations for all time steps. Of course, varying delta
formations result in a variation in backwater effects. When both bedload and suspended-load are considered, the backwater
effects are more dramatic.

The delta formed after 30 years are shown in Fig. 6.38A and B. After 30 years, the total load condition produces a delta reaching
a length of approximately 36 km. Considering only bedload results in a prediction of a delta only 28.2 km long. Holding the
downstream elevation constant results in a considerable backwater effect driven by sedimentation. Although this elevation
assumption is not wholly realistic, the results serve to illustrate the threat of flooding associated with reservoir sedimentation.
As the reservoir “silts” in, it will be unable to hold the same amount of water without producing a commensurate increase in
water stage.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.38a Delta location and height after 30 years; (A) bedload only.

Figure 6.38b Delta location and height after 30 years.

If one is interested in estimating the amount of time necessary to fill a reservoir, it is possible to simply divide the reservoirs
“filling” volume per unit width by the normal sediment inflow at the upstream end. To determine the “filling” volume, it is
necessary to consider the initial bed condition and the full bed conditions. The filling volume per unit width is estimated at
360,000 m3 /m. Assuming bedload only and dividing the filling volume by the normal bedload inflow results in an approximate
filling time of approximately 130 years. If total load is considered, an approximate filling time of 35 years is determined. This
agrees well with the results of the model; Fig. 6.38b shows that the total-load model predicts that the reservoir will be
approximately “full” around 40 years. Neither the above calculation nor the developed computer model considers the effect of
sediment compaction, which may play an important role in increasing the time required to fill a reservoir.

In general, the above results clearly indicate that suspended load plays a major role in reservoir sedimentation. Not considering
suspended-load results in a considerable underestimation of the progress and effects of reservoir sedimentation. If the
suspended load of the incoming flows is high, plunging may occur and turbidity currents will develop. Turbidity flows can
transport fine-grained sediment for long distances, hence having a profound effect on reservoir sedimentation and water
quality.

6.12. HYDRAULICS OF TURBIDITY CURRENTS


6.12.1. Introduction
© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
6.12.1. Introduction
Turbidity currents are currents of water laden with sediment that move downslope in otherwise still bodies of water. Consider
the situation illustration in Fig. 6.39. After plunging, a turbidity current moves along the bed of a laboratory reservoir (Bell, 1942).
It is seen that when the flow goes from the sloping portion onto the flat portion, there is a two-fold increase in current thickness,
indicating a change in flow regime through a hydraulic jump. There are a number of field situations where a similar slope-
induced hydraulic jump can take place (García, 1993, García, and Parker, 1989).

Figure 6.39 Turbidity current flowing into a laboratory reservoir. (Bell, 1942)

An important engineering aspect of turbidity currents concerns the impact these flows have on the water quality and
sedimentation in lakes and reservoirs. Turbidity flows were observed in lakes and man-made reservoirs long before their
occurrence in the ocean became apparent. This situation usually occurs during flood periods, when rivers carry a large amount
of sediment in suspension. In China, where the suspended load in most rivers is extremely large, the venting of turbidity
currents through bottom outlets to reduce the siltation of reservoirs has become common practice. Even though the bed slopes
of lakes and reservoirs are orders of magnitude smaller than those in the ocean, turbidity currents are still capable of traveling
long distances without losing their identities: e.g., more than 100 km in Lake Mead. An excellent account of numerical methods
to model turbidity currents in reservoirs can be found in Sloff (1997).

The ability of turbidity currents to transport sediment also has been put to use for the disposal of mining tailings (Normark and
Dickson, 1976) and ash from power station boilers. Environmental concern has reduced waste disposal into lakes, but in the
ocean, the dumping of mining tailing continues (Hay, 1987a, 1987b).

6.12.2. Governing Equations


A detailed derivation of the governing equations for two-dimensional turbidity currents can be found in Parker et al. (1986).
Here, the equations of motion are presented in layer-averaged form. The situation described in Fig. 6.40 is considered. A steady,
continuous turbidity current is flowing downslope through a quiescent body of water, which is assumed to be infinitely deep and
unstratified except for the turbidity current itself. The cross section is taken to be rectangular, with a width many times longer
than the underflow thickness; therefore, variation in the lateral direction can be neglected. The bed has a constant small slope S
and is covered with uniform sediment of geometric mean diameter Dsg and fall velocity vs; the x coordinate is directed
downslope tangential to the bed, and the z coordinate is directed upward normal to the bed. The submerged specific gravity of
the sediment is denoted by R = (ρs/ρ – 1), where ρs is the density of the sediment and ρ is the density of the clear water. Local
mean downstream-flow velocity and volumetric sediment concentration are denoted as u and c, respectively. The suspension is
dilute, hence c « 1 and Rc « 1 are assumed to hold everywhere. The parameters u and c are assumed to maintain similar profiles
as the current develops in the downslope direction. The layer-averaged current velocity U and volumetric concentration C and
the layer thickness h are defined via a set of moments (Parker et al., 1986):

(6.211a)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.211b)

(6.211c)

The equation of fluid mass balance integrates in the upward normal direction to yield

(6.212)

where ew is the coefficient of entrainment of water from the quiescent water above the current. The equation of sediment
conservation takes the layer-averaged form

(6.213)

where cb is the near-bed concentration of suspended sediment evaluated atz = 0.05h and Es is a dimensionless coefficient of
bed sediment entrainment into suspension. The integral momentum balance equation takes the form

(6.214)

where u* denotes the bed-shear velocity. The equations of sediment mass, fluid mass, and flow momentum balance must be
closed appropriately with algebraic laws for ew, u*, Es, and cb. The water entrainment coefficient ew is known to be a function of
the bulk Richardson number (Ri), which can be defined as

(6.215)

Figure 6.40 Turbidity current flowing downslope through a quiescent body of water. (after García,
1994)

and is equal to one over the square of the densimetric Fr = U/(gRCh) 1/2, often used in stratified flow studies. A useful equation
for the water entrainment coefficient plotted in Fig. 6.41 is the following (Parker et al., 1987):

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
(6.216)

Figure 6.41 Water entrainment coefficient as a function of Richardson number (After Parker et al.,
1987)

It is customary to take the bed shear stress to be proportional to the square of the flow velocity so thatu*2 = CDU2 , where CD is a
bed friction coefficient. Values of CD for turbidity currents have been found to vary between 0.002 and 0.05, as shown in Fig.
6.42 (Parker et al., 1987). The near-bed concentration cb can be related to the layer-averaged concentration C by a shape factor
ro = cb/C, which is approximately equal to 2 for sediment-laden underflows, as shown in Fig. 6.43 (Parker et al., 1987). The
sediment entrainment coefficient Es is known to be a function of bed shear stress and sediment characteristics (García and
Parker, 1991). The formulation of García and Parker (Eq. 6.173a) is plotted in Fig. 6.44, where data on sediment entrainment by
sediment-laden density currents also are included (García and Parker, 1993).

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.42 Plot of bed friction coefficient cD versus Reynolds number. (after Parker et al., 1987)

Figure 6.43 Plot of shape factor ro versus μ = u*/Vs. (after Parker et al., 1987)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.44 Plot of the sediment entrainment coefficient Es for both open-channel suspensions and
density currents. (after García and Parker, 1993)

6.12.3. Plunging Flow


The necessary conditions for plunging to occur in a reservoir may vary as a function of the physical parameters that produce
flow stratification. These parameters are sometimes known in advance from measurements in the field. Akiyama and Stefan
(1984) generalized several expressions that were derived from laboratory experiments, field measurements, or theoretical
analysis as a function of the parameters involved in plunging:

(6.217)

2 1/3

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
Figure 6.45 Plot of plunging flow depth versus (q2 /g′)1/3, including field and laboratory data García
(1996)

where hp = flow depth at plunging, qw = flow discharge per unit width, and Frp = densimetric Fr at plunging defined by

(6.218)

The value of Fr p has been found to range from 0.2 to 0.8 (Morris and Fan, 1998). If there is not enough suspended sediment,
plunging will not occur and a turbidity current will not develop. Figure 6.45 shows field and laboratory data for the flow depth at
plunging as a function of the inflow parameters (García, 1996).

6.12.4. Internal Hydraulic Jump


The bulk Ri, given by Eq. (6.215), is an important parameter governing the behavior of stratified slender flows, such as turbidity
currents (Turner, 1973). This parameter has a critical value Ric near unity so that the range Ri < Ric corresponds to a high-
velocity supercritical turbid flow regime, and the range Ri > Ric corresponds to a low-velocity subcritical turbid flow regime. The
change from supercritical flow to subcritical flow is accomplished via an internal hydraulic jump, as illustrated in Fig. 6.39.
Therein, a turbidity current undergoes a hydraulic jump induced by a change in bed slope in the proximity of a laboratory
reservoir. Conservation of momentum gives the following relation (García, 1993):

(6.219)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
which is analogous to Belanger’s equation for open-channel flow hydraulic jumps. For a known prejump Ri, 1 Eq. (6.219) gives
the ratio of the sequent current thickness h2 to the initial current thickness h1 . The subcritical flow, forced by some type of
control acting farther downstream, will influence the location of the jump and thus the length of the water-entraining
supercritical flow upstream of the jump. In laboratory experiments, the downstream boundary conditions are usually imposed
by the experimenter (e.g., weir, sluice gate, outfall) because of the finite length of experimental facilities. In the ocean or lakes,
where a current may travel several hundred kilometers without losing its identity, the control of the flow will operate through
deposition of sediment and bed friction.

6.12.5. Application: Turbidity Current in Lake Superior


As an example, the case of turbidity currents produced by the discharge of taconite tailings by the Reserve Mining Company into
Lake Superior at Silver Bay, Minnesota, is considered. Over a period of 20 years, the man-made turbidity currents formed a delta
with a steep front followed by a depositional fan. Normark and Dickson (1976) used field observations to infer that the
transition from the delta slope to the fan slope took place through a hydraulic jump, whereas Akiyama and Stefan (1985) used
numerical modeling to show a clear tendency by the flow to become supercritical shortly after reaching the fan region. However,
the lack of knowledge about the role played by the hydraulic jump has made flow computations along the subcritical region
practically impossible. Such computations can now be simplified through the knowledge gained in laboratory experiments
(García, 1994). This is illustrated by the following numerical experiment.

Figure 6.46 Simulation of turbidity current undergoing a hydraulic jump in Lake Superior,
Minnesota. (after García, 1993)

The lake bed topography at Silver Bay in Lake Superior is modeled in a one-dimensional configuration, as illustrated inFig.
6.46. The delta slop angle is 17º, and the fan slope is 1.5º. The delta-fan slope transition takes place between 600 m to 900 m
from the shore. The equations of motion (6.212), (6.213), and (6.214) are solved using a simple standard step method (García
and Parker, 1986). The water entrainment ew and sediment entrainment Es coefficients are estimated with relationships
proposed by Parker et al. (1987) and García and Parker (1991; 1993), respectively. A constant bed friction coefficient CD = 0.02
and a shape factor ro = 2 are used. Initial flow conditions at the tailings discharge point similar to those used by Akiyama and
Fukushima (1986) are used: i.e., Uo = 0.6 m/s, ho = 1 m, Φo = 0.1 m2 /s, and Rio = 0.5. The tailings have a mean particle size Dsg =
40 mm and a submerged specific gravity R = 2.1. Particle fall velocity is estimated to be vs = 0.14 cm/s. Lateral spreading of the
flows is ignored. The computations march downslope starting at the head of the delta, and at approximately 0.6 km from the
tailings’ discharge point, the flow starts to slow down because of the slope transition. If the current depth at the end of the fan
region could be known, the jump location could be determined with a simple “backwater” computation. Because this
information is not available, the hydraulic jump is assumed to take place at 0.9 km from the inlet. According to the laboratory
observations, water entrainment from above, as well as bed sediment entrainment into suspension, can be neglected after the
jump. Under these assumptions, Eq. (6.213) can be integrated with the help of Eq. (6.212), and an expression for the spatial
variation of the volumetric layer-averaged sediment concentration C is obtained,

(6.220)

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
where C is the value of Cj at the hydraulic jump and x′ is distance measured from the jump’s location. Since the flow discharge
per unit width qw is constant in the subcritical flow region (ew 0), Eq. (6.220) can be used to compute the volumetric sediment
transport rate per unit width CUh, at any location after the jump. The variation in current thickness between the jump’s location
and a point located 2.2 km from the inlet is shown in Fig. 6.46. The profile is obtained by first computing the value of C at 2.2
km with the help of Eq. (6.220), then by doing a “backwater” computation in an iterative manner until the computed current
thickness at 0.9 km coincides with the current thickness obtained with the supercritical flow computation and the hydraulic
jump Eq. (6.219). The flow discharge per unit width computed at the jump’s location is qw = 48 m2 /s. For such flow discharge,
Eq. (6.220) predicts that the turbidity current, after experiencing a hydraulic jump, will travel approximately 80 km before dying
out as a result of deposition of sediment.

6.13. REFERENCES
1. Akiyama and Fukushima, 1986, “Entrainment of noncohesive sediment into suspension.” 3rd Int.Symp. on River
Sedimentation, S. Y. Wang, H. W. Shen, and L. Z. Ding, eds., Univ. of Mississipi, 804–813.
2. Akiyama and Stefan, 1984, “ Plunging Flow into a Reservior: Theory,” (American Society of Civil Engineer. Journal of
Hydraulic Engineering) 110(4), 484–499
3. Akiyama and Stefan, 1985, “ Turbidity Current with Erosion and Deposition,” (American Society of Civil Engineering),
Journal of Hydraulic Engineering. 111(12), 1473–1496.
4. Ackers, P., and W. R. White, “Sediment Transport: New Approach and Analysis,” Journal of Hydraulic Engineering,
American Society of Civil Engineer99:2041–2060, 1973.
5. Allen, J. R. L., Physical Processes of Sedimentation, Elsevier, New York, 1970.
6. Amsler, M. L., and M. H. García, Discussion of “Sand-Dune Geometry of Large Rivers During Floods,” Journal of Hydraulic
Engineering, American Society of Civil Engineer123:582–584, 1997.
7. Anderson, A. G., “The Characteristics of Sediment Waves Formed by Flow in Open Channels,”Proceedings of the 3rd
Midwestern Conference on Fluid Mechanics, University. of Minnesota, Minneapolis, March 1953.
8. Annambhotla, V. S., W. W. Sayre, and R. H. Livesey, “Statistical Properties of Missouri River Bed Forms,” Journal
Waterways, Harbors, & Coastal Engineering Division, American Society of Civil Engineer98(WW4):489–510, 1972.
9. Ashida, K., and M. Michiue, “Study on Hydraulic Resistance and Bed-load Transport Rate in Alluvial Streams.”
Transactions of the Japan Society of Civil Engineering206:59–69, 1972.
10. Ashworth, P .J., S. J. Bennett, J. L. Best, and S. J. Mclelland, eds., Coherent Flow Structures in Open Channels, John Wiley
& Sons, Chichester, England, 1996.
11. Bell, H. S., Stratified Flow in Reservoirs and Its Use in Prevention of Silting, Miscellaneous PublicationNo. 491, U.S.
Department of Agriculture, 1942.
12. Bhowmik, N. G., A. P. Bonini, W. C. Bogner, and R. P. Byrne, Hydraulics of Flow and Sediment Transport in the Kankakee
River in Illinois, Report of Investigation 98, Illinois State Water Survey, Champaign, IL, 1980.
13. Blondeaux, P., G. Seminara, and M. Tubino, “A Unified Bar-Bend Theory of River Meanders,”Journal of Fluid Mechanics,
157:449–470, 1985.
14. Bogardi, J., Sediment Transport in Alluvial Streams, Akademiai Kiado, Budapest, Hungary, 1974.
15. Bouvard, M., Mobile Barrages and Intakes on Sediment Transporting Rivers, International Association for Hydraulic
Research Monograph Series, A. A. Balkema, Rotterdam, 1992.
16. Bridge, J. S., and S. J. Bennett, “A Model for the Entrainment and Transport of Sediment Grains of Mixed Sizes, Shapes,
and Densities,” Water Resources Research, 28:337–363, 1992.
17. Brooks, N. H., “Calculation of Suspended Load Discharge from Velocity and Concentration Parameters,” Proceedings of
the Federal Interagency Sedimentation Conference, Miscellaneous Publication No. 970, Agricultural Research Service,
1963, pp. 229–237.
18. Brownlie, W. R., “ Prediction of Flow Depth and Sediment Discharge in Open Channels,” Report No. KH–R–43A, Keck
Laboratory of Hydraulics and Water Resources, California Institute of Technology, Pasadena, CA. 1981a.
19. Brownlie, W. R., “Re-examination of Nikuradse Roughness Data,” Journal of the Hydraulics Division, American Society of
Civil Engineer 107(HY1):115–119, 1981b.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
20. Brownlie, W. R., Errata, “Re-examination of Kikuradse Roughness Data,” Journal of the Hydraulics Division, American
Society of Civil Engineer 108(HY8):971, 1982.
21. Carling, P. A., and M. R. Dawson, Advances in Fluvial Dynamics and Stratigraphy, John Wiley & Sons, Chichester, England,
1996.
22. Celik, I., and W. Rodi, “Modeling Suspended Sediment Transport in Nonequilibrium Situations,”Journal of Hydraulic
Engineering, American Society of Civil Engineer114:1157–1191, 1988.
23. Chabert, J., and J. L. Chauvin, “Formation de Dunes et de Rides dans les Modèles Fluviaux,”Bull. Cen. Rech. ess. Chatou,
No. 4, 1963.
24. Chang, H. H., Fluvial Processes in River Engineering, John & Sons Wiley, New York, 1988.
25. Coleman, S. E., and B. W. Melville, “Bed-Form Development,”Journal of Hydraulics Engineering, American Society of Civil
Engineer120:554–560, 1994.
26. Colombini, M., G. Seminara, and M. Tubino, “Finite-Amplitude Bars,”Journal of Fluid Mechanics, 181:213–232, 1987.
27. Colombini, M., M. Tubino, and P. Whiting, “Topographic Expression of Bars in Meandering Channels,” Proceedings of the
3rd International Workshop on Gravel-Bed Rivers, Florence, Italy, 1990.
28. Coussot, P., Mudflow Rheology and Dynamics, International Association for Hydraulic Research
29. Monograph Series, A.A. Balkema, Rotterdam, Netherlands, 1997.
30. Dietrich, W. E., “Settling Velocities of Natural Particles,” Water Resources Research, 18:1615–1626, 1982.
31. Diplas, P., and G. Parker, Pollution of Gravel Spawning Grounds Due to Fine Sediment, Report No. 240, EPA/R-808683-01-1,
St. Anthony Falls Hydraulics Laboratory, Univ. of Minnesota, Minneapolis, Minnesota. 1985.
32. Diplas, P., and A. Sutherland, “Sampling Techniques for Gravel Sized Sediments,” American Society of Civil
EngineerJournal of Hydraulic Engineering, 114:481–501, 1988.
33. Einstein, H. A., The Bedload Function for Bedload Transportation in Open Channel Flows, Technical Bulletin No. 1026,
U.S.D.A., Soil Conservation Service, 1950.
34. Einstein, H. A., and N. L. Barbarossa, “River Channel Roughness,” Transactions of the American Society of Civil Engineers,
117:1121–1146, 1952.
35. Engelund, F., “Instability of Erodible Beds,’ Journal of Fluid Mechanics, 42:225–244, 1970.
36. Engelund, F., and J. Fredsoe, “A Sediment Transport Model for Straight Alluvial Channels,” Nordic Hydrology, 7:293–306,
1976.
37. Engelund, F., and E. Hansen. A Monograph on Sediment Transport in Alluvial Streams, Teknisk Vorlag, Copenhagen,
Denmark, 1967.
38. Exner, F. M., “Über die Wechselwirkung Zwischen Wasser und Geschiebe in Flüssen,” Sitzenberichte der Academie der
Wissenschaften, Wien, Austria, Sec. IIA, 134:199, 1925.
39. Fernandez Luque, R., and R. van Beek, “Erosion and Transport of Bed Sediment.” Journal of. Hydraulics Research, 14:127–
144, 1976.
40. Fredsoe, J., “On the Development of Dunes in Erodible Channels,”Journal of Fluid Mechanics, 64(1):1–16, 1974.
41. Fredsoe, J., “Shape and Dimensions of Stationary Dunes in Rivers,” Journal of the Hydraulics Division,American Society of
Civil Engineer108(HY8):932–947, 1982.
42. Fredsoe, J., and R. Deigaard, Mechanics of Coastal Sediment Transport, Advances Series on Ocean Engineering, Vol. 3,
World Scientific Publishing, Singapore, 1992.
43. García, M. H., “Hydraulic Jumps in Sediment-Laden Bottom Currents,” Journal of Hydraulic Engineering, American Society
of Civil Engineer 199:1094-1117, 1993.
44. García, M. H., “Depositional Turbidity Currents Laden with Poorly-Sorted Sediment,” Journal of Hydraulic Engineering,
American Society of Civil Engineer 120:1240–1263, 1994.
45. García, M. H., Environmental Hydrodynamics, Universidad Nacional del Litoral Publication Center, Santa Fe, Argentina (in
Spanish), 1996.
46. García, M. H., and Y. Niño “Lagrangian Description of Bedload Transport by Saltating Particles,”Proceedings of the 6th
International Symposium Stochastic Hydraulics,Taipei, Taiwan, 259-266, 1992.
47. García, M. H., and Y. Niño, “Dynamics of Sediment Bars in Straight and Meandering Channels: Experiments on the
Resonance Phenomenon,” IAHR Journal of Hydraulic Research, 31:739–761, 1993.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
48. García, M. H., and G. Parker, “On the Numerical Prediction of Turbidity Currents,” Proc. 3rd Int’l Symp. River Sedimentation,
The University of Mississippi, University, Mississippi, 1556-1565, 1986.
49. García, M. H. and G. Parker, “Experiments on Hydraulic Jumps in Turbidity Currents Near a Canyon-Fan Transition,”
Science, 117:393–396, 1989.
50. García, M., andG. Parker, “Entrainment of Bed Sediment into Suspension,” Journal of Hydraulic Engineering, American
Society of Civil Engineer 117:414–435, 1991.
51. García, M., and G. Parker, “Experiments on the Entrainment of Sediment into Suspension by a Dense Bottom Current,”
Journal of Geophysical Research, 98(C3):4793–4807, 1993.
52. Garde, R. J., and Albertson, 1959. “Sand Waves and Regimes of Flow in Alluvial Channels”Proceedings of the 8th
International Association for Hydraulic Research Congress, Montreal, Canada, Vol. IV, pa.per 28
53. Garde, R. J., and K. G. Ranga Raju, Mechanics of Sediment Transportation and Alluvial Stream
54. Problems, 2nd ed., Wiley Easter Limited, New Delhi, India, 1985.
55. Gill, M. A., “Height of Sand Dunes in Open Channel Flows,”Journal of the Hydraulics Division, American Society of Civil
Engineer97(HY12):2067–2074, 1971.
56. Graf, W. H., Hydraulics of Sediment Transport, McGraw-Hill, New York, 1971.
57. Haque, M. I., and K. Mahmoud, “Geometry of Ripples and Dunes,” Journal of Hydraulic Engineering, American Society of
Civil Engineer111(1):48–63, 1985.
58. Hay, A. E., “Turbidity Currents and Submarine Channel Formation in Rupert Inlet, British Columbia, 1: Surge Observations,”
Journal of Geophysical Research, 92(C3):2875–2881, 1987a.
59. Hay, A. E., “Turbidity Currents and Submarine Channel Formation in Rupert Inlet, British Columbia, 2: The Roles of
Continuous and Sure-Type Flow,” Journal of Geophysical Research, 92(C3):2883–2900, 1987b.
60. Hayashi, T., “Formation of Dunes and Antidunes in Open Channels,” Journal of the Hydraulics Division, American Society
of Civil Engineer96(HY2):357–367, 1970.
61. Hino, M., “Equilibrium-Range Spectra of Sand Waves Formed by Flowing Water,”Journal of Fluid Mechanics, 34(3):565–
573, 1968.
62. Hotchkiss, R. H., and G. Parker, “Shockfitting of Aggradational Profiles Due to Backwater,” Journal of Hydraulic
Engineering, American Society of Civil Engineer 117(9):1129–1144, 1991.
63. Ikeda, S., “Incipient Motion of Sand Particles on Side Slopes,” Journal of the Hydraulics Division, American Society of Civil
Engineer108(HY1):95–114, 1982.
64. Ikeda, S., “Prediction of Alternate Bar Wavelength and Height,” Journal of the Hydraulics Division, American Society of
Civil Engineer110(4):371–386, 1984.
65. Ikeda, S., and G. Parker, eds., “River Meandering,” Water Resources Monograph No. 12, American Geophysical Union,
Washington D.C1989.
66. Jaeggi, M. N. R., “Formation and Effects of Alternate Bars,” Journal of Hydraulic Engineering, American Society of Civil
Engineer10:142–156, 1984.
67. Jain, S. C., and J. F. Kennedy, “The Spectral Evolution of Sedimentary Bed Forms,”Journal of Fluid Mechanics, 63:301–
314, 1974.
68. Jansen, P. P., L. van Bendegom, J. van den Berg, M. de Vries, and A. Zanen, Principles of River Engineering: The Non-Tidal
Alluvial River, Pitman, London, UK, 1979.
69. Julien, P.Y., Erosion and Sedimentation, Cambridge University Press, 1995.
70. Julien, P. Y., and G. J. Klaassen, “Sand-Dune Geometry of Large Rivers during Floods,”Journal of Hydraulic Engineering,
American Society of Civil Engineer121:657–663, 1995.
71. Kennedy, J. F., “The Mechanics of Dunes and Antidunes in Erodible-Bed Channels,”Journal of Fluid Mechanics, 16:521-
544, 1963.
72. Kennedy, J. F., 1969. “The Formation of Sediment Ripples, Dunes, Antidunes,” Annual R2 views of Fluid MechanicsVol. 1,
pp. 147–167, 1969.
73. Kinoshita, R., and H. Miwa, “River Channel Formation Which Prevents Downstream Translation of Transverse Bars” (in
Japanese), Shinsabo, 94:12–171974.
74. Klaassen, G. J., H. J. M. Ogink, and L. C. van Rijn, “DHL-Research on Bed Forms, Resistance to Flow and Sediment
Transport,” Procedings of the 3rd International Symposium River Sedimentation, Jackson, MI, March-April 1986, PP. 58–

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
820
75. Lane, E.W., “Design of Stable Channels,” Transactions of the American Society of Civil Engineer, 81(745):1–17, 1955.
76. Lavelle, J. W., and H. O. Mofjeld, “Do Critical Stresses for Incipient Motion and Erosion Really Exit?”ASCE Journal of
Hydraulic Engineering, 113:370–385, 1987.
77. Liu, H. K., “Mechanics of Sediment-Ripple Formation,” American Society of Civil Engineer Journal of the Hydraulics
Division, 83(HY2):1–23, 1957.
78. Menduni, G., and E. Paris, AOn the Prediction of Geometric Characteristics of Dunes,” Proceedings of the International
Symposium on River Sedimentation, Jackson, MI, March-April 1986, pp. 665–674.
79. Mehta, A. J., E. J. Hayter, W. R. Parker, R. B. Krone, and A. M. Teeter, “Cohesive Sediment Transport I: Process
Description,” Journal of Hydraulic Engineering, American Society of Civil Engineer115:1076–1093, 1989a.
80. Mehta, A. J., E. J. Hayter, W. R. Parker, R. B. Krone, and A. M. Teeter, “Cohesive Sediment Transport II:
Application,”Journal of Hydraulic Engineering, American Society of Civil Engineer115:1094–1113, 1989b.
81. Meyer-Peter, E., and R. Muller, “Formulas for Bedload Transport.”Proceedings of the 2nd Congress of the International
Association for Hydraulic Research, Stockholm, 1948, pp. 39–64,
82. Milhous, R. T., “Sediment Transport in a Gravel Bottomed Stream,” doctoral dissertation, Oregon State University, Corvallis,
1973.
83. Misri, R. I., Garde, R. J., and Ranga Raju, K. G., 1983, “Experiments on Bedload Transport of Nonuniform sands and
Gravels,” in Proceedings of 2nd International Symposium on River Sedimentation, Water Resources and Electric power
Press, China, 450-450.
84. Morris, G. L., and J. Fan, Reservoir Sedimentation Handbook, Mc Graw–Hill Book Co, N.Y., 1988
85. Nakagawa, H., and T. Tsujimoto, “Special Analysis of San Bed Instability,” Journal of Hydraulic Engineering, American
Society of Civil Engineer110:457–483, 1984.
86. Nakato, T., and R. Ettema, eds., Issues and Directions in Hydraulics, Proceedings of the Iowa Hydraulics College, Iowa
City, May 1995, A. A. Balkema, Rotterdam, 1996.
87. National Research Council, “Alluvial Fan Flooding,” Water Science and Technology Board, National Academy Press,
Washington, DC, 1996.
88. Nelson, J. M., and J. D. Smith, “Flow in Meandering Channels with Natural Topography,” in S. Ikeda and G. Parker, eds.,
River Meandering, Water Resources Monograph No. 12, American Geophysical Union, 1989, pp. 69–102.
89. Nielsen, P., Coastal Bottom Boundary Layers and Sediment Transport, Advances Series on Ocean Engineering, Vol. 4,
World Scientific Publishing Singapore, 1992.
90. Nikuradse, J., “Laws of Flow in Rough Pipes” (in German), National Advisory Committee for Aeronautics Tech Memo No.
1292, Washington, DC, 1950.
91. Niño, Y., and M. H. García, “Gravel Saltation II: Modeling,” Water Resources Research, 30(6):1915–1924, 1994.
92. Niño, Y., and M. H. García, “Using Langrangian Particle Saltation Observations for Bedload Sediment Transport Modelling,”
Hydrological Processes, 12:1197–1218, 1998.
93. Niño, Y., M. H. García, and L. Ayala, “Gravel Saltation I: Experiments,” Water Resources Research, AGU 30:1907–1914,
1994.
94. Nordin, C. F. “ Aspects of flow resistance and sediment transport: Rio Grande ner Bernalillo,” Water supply paper No.
1498-H, U.S. Geological Survey, Washington, D C1963.
95. Nordin, C. F. Jr., and J. H. Algert, “Spectral Analysis of Sand Waves,” Journal of the Hydraulics Division, American Society
of Civil Engineer92(HY5):95–114, 1966.
96. Normark and Dickson, 1976, “Man–mode Turbidity Currents in Lake Superior,” Sedimentology, Vol. 23, pp 815–831.
97. Parker, G., “Sediment Inertia as a Cause of River Antidunes,” Journal of the Hydraulics Division, American Society of Civil
Engineer101(HY2):211–221, 1975.
98. Parker, G., “Self-Formed Straight Rivers with Equilibrium Banks and Mobile Bed, Part 2. The Gravel River,” Journal of Fluid
Mechanics, 89(1):127–146, 1978.
99. Parker, G., “Surface-Based Bedload Transport Relation for Gravel Rivers,” Journal of Hydraulic Research, 28(4):417–436,
1990.
100. Parker, G., “Selective Sorting and Abrasion of River Gravel: Theory,” Journal of Hydraulic Engineering, American Society of
Civil Engineer117(HY2):131–149, 1991a.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
101. Parker, G., “Selective Sorting and Abrasion of River Gravel: Applications,” Journal of Hydraulic Engineering, American
Society of Civil Engineer117:(HY2):150–171, 1991b.
102. Parker, G., and A. G. Anderson, “Basic Principles of River Hydraulics,” Engineer Journal of the Hydraulics Division,
American Society of Civil Engineer103(HY9):1077–1087, 1977.
103. Parker, G., and E. D. Andrews, “Sorting of Bedload Sediment by Flow in Meander Bends,”Water Resources Research,
24:1361–1373, 1985.
104. Parker and Ikeda, 1989, “River Meandering,” Water Resources Mouograph Nº 12, American Geophysical Union,
Washington D.C. 485p.
105. Parker, G., and A. J. Sutherland, “Fluvial Armor,” IAHR Journal of Hydraulic Research, 28:529–544, 1990.
106. Parker, G., M. García, Y. Fukushima, and W. Yu, “Experiments on Turbidity Currents over an Erodible Bed,” Journal of
Hydraulic Research, 25(1):123–147, 1987.
107. Press, W. H., B. P. Flannery, S. A. Teulovsky, and W. T. Vetterling, Numerical Recipes, Cambridge University Press, New
York, 1986.
108. Ranga Raju, K. G., and Soni, J. P., “Geometry of Ripples and Dunes in Alluvial Channels,” IAHR Journal of Hydraulic
Reserach, 14:241–249, 1976.
109. Raudkivi, A. J., Loose Boundary Hydraulics, 3rd ed., Pergamon Press, New York, 1990.
110. Raudkivi, A. J., “Sedimentation: Exclusion and Removal of Sediment fro Diverted Water,”Hydraulic Structures Design
Manual, International Association for Hydraulic Research, A. A. Balkema, Rotterdam, 1993.
111. Raudkivi, A. J., and H. H. Witte, “Development of Bed Features,” Journal of Hydraulic Engineering, American Society of
Civil Engineer 116:1063–1079, 1990.
112. Renard, K. G., G. R. Foster, G. S. Weesies, D. K. McCool, and D. C. Yoders, coordinators,Predicting Soil Erosion by Water: A
Guide to Conservation Planning with the Revised Universal Soil Loss Equation (RUSLE), Agriculture Handbook No. 703,
U.S. Department of Agriculture, Washington, DC, 1997.
113. Reynolds, A. J., “Waves on the Bed of an Erodible Channel,”Journal of Fluid Mechanics, 22(1):113-133, 1965.
114. Richards, K. J., “The Formation of Ripples and Dunes on Erodible Bed,”Journal of Fluid Mechanics, 99:597–618, 1980.
115. Rouse, H., “Modern Conceptions of the Mechanics of Turbulence,” Transactions of the American Society of Civil
Engineers,102, 1957.
116. Schlichting, H., Boundary Layer Theory, 7th ed., McGraw-Hill, New York, 1979.
117. Sekine, M., and H. Kikkawa, “Mechanics of Saltating Grains II,” Journal of Hydraulic Engineering, American Society of Civil
Engineer118:536–558, 1992.
118. Sekine, M., and G. Parker, “Bedload Transport on a Transverse Slope I”Journal of Hydraulic Engineering, American
Society of Civil Engineer118:513–535, 1992.
119. Shields, A., “Anwendung der Aenlichkeitsmechanik und der Turbulenzforschung auf die Geschiebebewegung,”
Mitteilungen der Preussischen Versuchsanstalt fur Wasserbau und Schiffbau, Berlin, Germany by W. P. Ott and J. C. van
Uchelen, trans., California Institute of Technology, Pasadena, CA, 1936.
120. Sieben, J., “Modelling of Hydraulics and Morphology in Mountain Rivers,” doctoral dissertation, Delft University of
Technology, Netherlands, 1997.
121. Simons, D. B., and E. V. Richardson, “Resistance to Flow in Alluvial Channels,”Professional Paper No. 422J, U.S.
Geological Survey, 1966.
122. Simons, D. B., and F. Senturk, Sediment Transport Technology, rev. ed., Waters Resources Publications, Littleton, Co,
1992.
123. Sloff, C. J., “Sedimentation in Reservoirs,” doctoral dissertation, Delft University of Technology, Netherlands, 1997.
124. Smith, J. D., “Stability of a Sand Bed Subjected to a Shear Flow of Low Froude Number,”J. Geophysical Research,
75(30):5928-5940, 1970.
125. Smith, J. D., and S. R. McLean, “Spatially Averaged Flow over a Wavy Surface,” Journal of Geophysical Research,
83:1735–1746, 1977.
126. Takahashi, T., Debris Flow, International Association for Hydraulic Research Monograph Series, A.A. Balkemabuoyancy
Rotterdam, 1991.
127. Taylor, B. D., and V. A. Vanoni, “Temperature Effects in Low-Transport, Flat-Bed Flows,” American Society of Civil Engineer
Journal of the Hydraulic Division, 98(HY12):2191-2206, 1972.

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.
128. Tubino, M., and G. Seminara, “Free-Forced Interactions in Developing Meanders and Suppression of Free Bars,”Journal of
Fluid Mechanics, 214:131-159, 1990.
129. Turner, 1973, “ Bouyancy Effects in Fluids,” Cambridge University Press, Cambridge, 368p.
130. van Rijn, L. C., “Sediment Transport, Part II. Suspended Load Transport,”Journal of Hydraulic
131. Engineering, 110:1613–1641, 1984a.
132. van Rijn, L. C., “Sediment Transport, Part III. Bed Forms and Alluvial Roughness,” American Society of Civil Engineer
Journal of Hydraulic Engineering, 110(12):1733-1754, 1984b
133. van Rijn, L. C., “Mathematical Modelling of Morphological Processes in the Case of Suspended Sediment Transport,” Delft
Hydraulics Communication No. 382, Delft University of Technology, Netherlands, 1987.
134. Vanoni, V. A. “Factors Determining Bed Forms of Alluvial Streams,” American Society of Civil EngineerJournal of the
Hydraulic Division, 100(HY3):363–377, 1974.
135. Vanoni, V. A., ed., Sedimentation Engineering, American Society of Civil Engineer Manuals and Reports on Engineering
Practice No. 54, American Society of Civil Engineers, New York, 1975.
136. Wan, Z., and Z. Wang, Hyperconcentrated Flow, International Association for Hydraulic Research Monograph Series, A.A.
Balkema, Rotterdam, 1994.
137. White, W. R., H. Milli, and A. D. Crabbe, “Sediment Transport I and II: An Appraisal of Available Methods,” HRS INT 119,
Hydraulics Research Station, Wallingford, UK, 1973.
138. Wiberg, P. L., and J. D. Smith, “Model for Calculating Bedload Transport of Sediment,” American Society of Civil Engineer
Journal of Hydraulic Engineering, 115(1):101–123, 1989.
139. Wiberg, P., and J. D. Smith, “Calculations of the Critical Shear Stress for Motion of Uniform and Heterogeneous
Sediments,” Water Resources Research, 23:1471–1480, 1987.
140. Wiberg, P. L., and J. D. Smith, “A Theoretical Model for Saltating Grains,” Journal of Geophysical Research, 90(C4):7341–
7354, 1985.
141. Wilcock, P. R., “Methods for Estimating the Critical Shear Stress of Individual Fractions in Mixed-Size Sediment,”Water
Resources Research, 24:1127–1135, 1988.
142. Wilson, K. C., “Bedload Transport at High Shear Stresses,” American Society of Civil EngineerJournal of Hydraulic
Engineering, 92(HY6):49–59, 1966.
143. Yalin, M. S., “An Expression for Bedload Transportation.” American Society of Civil EngineerJournal of the Hydraulics
Division, 89(HY3):221–250, 1963.
144. Yalin, M.S., “Geometrical Properties of Sand Waves,” American Society of Civil EngineerJournal of the Hydraulics Division,
90(HY5):105–119, 1964.
145. Yalin, M. S., Mechanics of Sediment Transport, Pergamon, Braunschweig, Germany. 1972.
146. Yalin, M. S., River Mechanics, Pergamon Press, New York, 1992
147. Yalin, M. S., and E. Karahan, “Steepness of Sedimentary Dunes,” American Society of Civil Engineer Journal of the
Hydraulics Division, 105(HY4):381–392, 1979.
148. Yang, C. T. “Incipient Motion and Sediment Transport,” American Society of Civil Engineer Journal of the Hydraulic
Division, 99:1679–1704, 1973.
149. Yang, C. T., Sediment Transport: Theory and Practice, McGraw-HillNew York, 1996.
150. Yen, B. C., ed., Channel Flow Resistance: Centennial of Manning’s Formula, Water Resources Publications, Littleton, CO,
1992

© McGraw-Hill Education. All rights reserved. Any use is subject to the Terms of Use, Privacy Notice and copyright information.

You might also like