Вы находитесь на странице: 1из 33

5.

1 INTRODUCTION
The thermal, chemical, and biologic quality of water in rivers, lakes, reservoirs, and near
coastal areas is inseparable from a consideration of hydraulic engineering principles;
therefore, the term environmental hydraulics. In this chapter we discuss the basic princi-
ples of water and thermal budgets as well as mixing and dispersion.
5.2 WATER AND THERMAL BUDGETS
5.2.1 Water Budget
A water budget is a statement of the law of conservation of mass or
(change in storage) (input) (output) (5.1)
and the expressions of the water budget can range from simple to very complex. For exam-
ple, consider the lake or reservoir shown in Figure 5.1. For this situation, a generic water
budget could be written as follows:

d
d
S
t
s
(I
c
I
o
I
g
P
r
R
r
) (E
v
T
r
G
s
O
c
W) (5.2)
CHAPTER 5
ENVIRONMENTAL
HYDRAULICS
Richard H. French
Water Resources Center
Desert Research Institute
University and Community College System of Nevada
Reno, Nevada
Steven C. McCutcheon
Ecosystems Research Division
National Exposure Research Laboratory
U.S. Environmental Protection Agency
Athens, Georgia
James L. Martin
AScI Corporation
Athens, Georgia
5.1
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
Source: HYDRAULIC DESIGN HANDBOOK
where I
c
channel inflow rate, I
o
overland inflow rate, I
g
groundwater inflow rate,
P
r
precipitation rate, R
r
return flow rate, E
v
evaporation rate, T
r
transpiration
rate, G
s
groundwater seepage rate, O
c
channel outflow rate, W consumptive with-
drawal, and S
s
lake/reservoir storage rate at time t (volume).
The solution of Eq (5.2) quantifies the terms, and, in many cases, the goal of the mod-
eling effort is to estimate the value of a single term or group of terms: for example, evap-
otranspiration (E
v
T
r
). The reliability of using a water budget is directly related to the
accuracy of the prediction techniques used, the availability and quality of gauged data, and
the time period involved. Among the methods of evaluating the individual terms in Eq.
(5.2) are the following:
Channel inflow and outflow ( I
c
and O
c
)gauging, statistical simulation.
Overland inflow (I
o
)gauging, rainfall-runoff relationships.
Groundwater inflow and seepage rate (I
g
and G
s
)seepage equations, gauging.
Precipitation (P
r
)gauging, statistical simulation (Smith, 1993).
Evaporation and transpiration (E and T)gauging, evaporation/transpiration predic-
tion relationships (Bowie et al. 1985; Shuttleworth, 1993).
Return flow and withdrawal (R
r
and W)gauging.
5.2 Chapter Five
FIGURE 5.1 A hypothetical lake illustrating the variables in the water budget.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.2.2 Thermal Budget
The total thermal budget for a body of water includes atmospheric heat exchange at the
air water interface (usually the dominant process), the effects of inflows (tributaries,
wastewater, and cooling water discharges), heat resulting from chemical-biological reac-
tions, and heat exchange with the stream bed. In the following sections, the primary com-
ponents of the air-water interface heat budget will be briefly discussed; for further details
the reader is referred to Bowie et al., (1985), McCutcheon (1989), or Shuttleworth
(1993).
Atmospheric heat exchange at the air-water interface is given by
H Q
s
Q
sr
Q
a
Q
ar
Q
br
Q
e
Q
c
(5.3)
where H net surface heat flux, Q
s
shortwave radiation incident to the water surface
[3300 (kcal/m
2
)/h], Q
sr
reflected shortwave radiation [525 (kcalm
2
)/h], Q
a
incom-
ing longwave radiation from the atmosphere (225360 kcal/m
2
/h), Q
ar
reflected long-
wave radiation [515 (kcalm
2
)/hr], Q
br
longwave back radiation emitted by the water
body [220345 (kcalm
2
)/h], Q
e
energy utilized by evaporation [25900 (kcalm
2
)/h],
and Q
c
energy convected to or from the body of water (3550 kcalm
2
/hr). Note that
the ranges given are typical for the middle latitudes of the United States (Bowie et al.,
1985).
The equations for estimating the terms of the thermal budgets use a mixed set of units,
and appropriate conversions among the different units used are provided in Table 5.1.
5.2.2.1 Net atmospheric shortwave radiation (Q
s
Q
sr
) The net shortwave radiation
(Q
sn
) is that portion of the incident shortwave radiation captured at the ground, taking into
account losses caused by reflection. Although solar radiation can be measured with spe-
cialized meteorological stations equipped with radiometers, these instruments require
painstaking calibration and maintenance. In most cases, measured values of solar radia-
tion are not available at the location of interest and must be estimated from equations.
Among the formulations for estimating net shortwave solar radiation is
Q
sn
Q
s
Q
sr
0.94Q
sc
(1 0.65C
2
c
) (5.4)
where Q
sc
clear sky solar radiation [kcalm
2
)/h) and C
c
fraction of sky covered by
clouds (Anderson, 1954; Ryan and Harleman, 1973). It is pertinent to note that Eq. (5.4)
Environmental Hydraulics 5.3
TABLE 5.1 Useful Energy Conversions for Energy Budget Calculations
1Btuft
2
/day = 0.131 W/m
2
= 0.271 Ly/day = 0.113 (kcalm
2
)/h
1 watt/m
2
= 7.61 Btuft
2
)/day = 2.07 Ly/day = 0.86 (kcalm
2
)/h
1 Ly/day = 0.483 W/m
2
= 3.69 (Btu/ft
2
)/day = 0.42 (kcalm
2
)/h
1 (kcalm
2
)/hr = 1.16 W/m
2
= 2.40 Ly/day = 8.85 (Btuft
2
)/day
1 kpa = 10 mb = 7.69 mm Hg = 0.303 in (Hg)
1 mb = 0.1 kpa = 0.769 mm Hg = 0.03 in (Hg)
1 mm Hg = 1.3 mb = 0.13 kpa = 0.039 in (Hg)
1 in Hg = 33.0 mb = 25.4 mm Hg = 3.3 kpa
Abbreviations Ly Langleys; mb millibar; and Btu British Thermal Unit
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.4 Chapter Five
assumes average reflectance at the Waters surface and uses clear sky solar radiation. In
some situations, the effects of atmospheric attenuation are much greater than normal and
more complex equations are required (e.g., 1972). Clear sky radiation (Q
sc
) can be esti-
mated as a function of calendar month and latitude from Fig. 5.2.
Shortwave solar radiation is absorbed at the waters surface and penetrates the water
column, depending on the wavelength of the radiation, the properties of the water, and the
matter suspended in the water. The degree of penetration of shortwave solar radiation
(sunlight) into the water column has a significant effect not only on water temperature but
also on the rate of photosynthesis by aquatic plants and the general clarity, color and aes-
thetic quality of the water. The penetration of shortwave solar radiation is described by
I I
o
exp (k
e
y) (5.5)
where I light intensity at depth y, K
e
extinction coefficient, and I
o
light intensity
at the surface (y 0).
Values of the extinction coefficient can be estimated by several methods. For example,
measurement of total light penetration into a water column can be made by using a pyre-
heliometer positioned at the surface that measures the total incoming solar radiation.
Simultaneously, an underwater photometer is lowered and the radiation is recorded at each
of a series of depths throughout the water column. Then, a value of K
e
can be estimated
by linear leastsquares regression. An alternative but traditional, simpler, and less accu-
rate method to estimate K
e
is to lower a target into the water column until, by eye, the tar-
get just disappears. A standardized target (Secchi disk) is commonly used, and a number
of investigators (Beeton, 1958; French et al., 1982; Sverdrup et al, 1942;) have developed
empirical relationships between, the Secchi disk depth (y
s
) and the extinction coefficient
of the form.
K
e

(
1.2 to
y
s
1.9)
(5.6)
Finally, the depth (y
e
) at which 1 percent of the surface radiation still remains (the
euphotic depth) is given from Eq. (5.5) as
y
e

4
K
.6
e
1
(5.7)
5.2.2.2 Net atmospheric long-wave radiation (Q
a
Q
ar
) Atmospheric radiation is char-
acterized by much longer wavelengths than solar radiation because the major emitting
elements are water vapor, carbon dioxide, and ozone. The approach generally used to
estimate this flux involves the empirical estimation of an overall atmospheric emissivity
and the use of the Stephan-Boltzman law (Ryan and Harleman, 1973). Swinbank (1963)
developed the following equation, which has been used in many water quality models:
Q
an
Q
a
Q
ar
1.16 10
13
(1 0.17C
2
c
)(T
a
460)
6
(5.8)
where Q
an
net longwave atmospheric radiation (Btuft
2
/day), C
c
fraction of sky cov-
ered by clouds, and T
a
dry bulb air temperature (F).
5.2.2.3 Long-wave back radiation (Q
br
) The long-wave back radiation from a water
surface in most cases is the largest of all the fluxes in the heat budget (Ryan
and Harleman, 1973). The emissivity of a water surface is well known; therefore, this
flux can be estimated with a high degree of accuracy as a function of the water surface
temperature:
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.5
F
I
G
U
R
E

5
.
2
C
l
e
a
r

s
k
y

s
o
l
a
r

r
a
d
i
a
t
i
o
n
.

(
F
r
o
m

H
a
m
o
n

e
t

a
l
.

1
9
5
4
)
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Q
br
0.97T
4
s
(5.9)
where Q
br
longwave back radiation (cal/m
2
/s), T
s
surface water temperature (
0
K),
and Stefan-Boltzman constant (1.357 10
8
calm
2
/s/K
4
)
5.2.2.4 Evaporative heat flux (Q
e
) Evaporative heat loss (kcal/m
2
/s) occurs as a result
of the change of state of water from a liquid form to vapor and is estimated by
Q
a
L
w
E
v
(5.10)
where L
w
latent heat of vaporization (597 0.57T
s
, kcal/kg), T
s
surface water tem-
perature (C), E
v
evaporation rate (m/s), and water density (kg/m
3
).
A standard expression for evaporation from a natural water surface is
E
v
(a bW)(e
s
e
a
) (5.11)
where E
v
evaporation rate (m/s), a and b empirical coefficients, W wind speed at
some specified distance above the water surface (m/s), e
s
saturation vapor pressure at
the temperature of the water surface (mb), and e
a
vapor pressure of the overlying atmos-
phere (mb). In many cases, the empirical coefficient a has been taken as zero with 1
10
9
b 5 10
9
(Bowie et al., 1985). The saturated vapor pressure can be estimated
(Thackston, 1974) by
e
s
exp
j
,
(
17.62

T
s
9

50
4
1
60

\
(
,
(5.12)
where e
s
is in inches of Hg, and T
s
water surface temperature (F). There are a number
of ways of estimating e
a
, depending on the available data. For example, if the relative
humidity (R
H
) is known, then
R
H

e
e
a
s
(5.13)
and then if the wet bulb temperature and atmospheric pressure are known (Brown and
Barnwell, 1987)
e
a
e
s
0.000367P
a
(T
a
T
wb
)
j
,
(
1

T
w
1
b
5

71
32

\
(
,
(5.14)
where all pressures are in (in Hg), all temperatures are in (F), P
a
atmospheric pressure,
and T
wb
wet bulb temperature. The relationship among the air and wet bulb tempera-
tures (F) and relative humidity (Thackston, 1974) is
T
wb
(0.655 0.36R
H
)T
a
(5.15)
There are many equations for estimating the rate of evaporation. For example, Jobson
(1980) developed a modified formula that was used in the temperature modeling of the
San Diego Aqueduct and subsequently was modified for use on the Chattahoochee River
in Georgia (Jobson and Keefer, 1979). McCutcheon (1982) noted that, in many models,
the wind speed function is a catchall term that compensates for many factors, such as (1)
numerical dispersion in some models, (2) the effects of wind direction, fetch, channel
width, sinuosity, bank, and tree height, (3) the effects of depth, turbulence, and lateral
velocity distribution; and (4) the stability of air moving over the stream. (Fetch is the dis-
tance over which the wind blows or causes shear over the waters surface.) Finally, it is
5.6 Chapter Five
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
important to note that evaporation estimators that work well for lakes or reservoirs will not
necessarily provide the same level of performance when used in streams, rivers, or con-
structed open channels.
5.2.2.5 Convective heat flux (Q
c
). Convective heat is transferred between air and water
by conduction and is transported to or from the air-water interface by convection. The con-
vective heat flux is related to the evaporative heat flux (Q
e
) by the Bowen ratio (Bowie et
al., 1985), or
R
B

Q
Q
c
e

(6.19 10
4
)P
a

T
e
s
s

T
e
a
a

(5.16)
where all temperatures are in (C), all pressures are in (mb), and R
B
Bowen ratio.
5.2.2.6 Conclusion. The foregoing is a brief summary of the approaches used most fre-
quently to estimate surface heat exchange in numerical models. The reader is referred to
other publications for a more detailed discussion of the approaches (Bowie et al., 1985)
and meteorological data requirements (Shanahan, 1984). Note that each situation should
be considered carefully from the viewpoint of specific factors that must be taken into
account. For example, in most lakes, estuaries, and deep rivers, the thermal flux through
the bottom is not significant. However, in water bodies with depths less than 3 m (10 ft),
bed conduction of heat can be significant in determining the diurnal variation of temper-
atures within the body of water (Jobson, 1980, Jobson and Keefer, 1979).
5.3 EFFECTS AND CAUSES OF STRATIFICATION
5.3.1 Effects
The density of water is strongly affected by temperature and the concentrations of dis-
solved and suspended solids. Regardless of the cause of differences in water density, water
with the greatest density is found at the bottom, whereas water with the least density resides
at the surface. When density gradients are strong, vertical mixing is inhibited. Stratification
is the establishment of distinct layers of water of different densities (Mills et al., 1982).
Stratification is enhanced by quiescent conditions and is destroyed by in a body of water-
phenomenasc that encourage mixing (wind stress, turbulence caused by large inflows, and
destabilizing changes in water temperature). In many bodies of water (rivers, lakes, and
reservoirs), stratification is the single most important phenomena affecting water quality.
When stratification is absent, the water column is mixed vertically and dissolved oxy-
gen (DO) is present in the vertical water column from the top to the bottom: that is, fully
mixed water columns do not have DO deficit problems. For example, when stratification
occurs, in reservoirs and lakes mixing is limited to the epliminion or surface layer. Since
stratification inhibits, vertical mixing is inhibited by stratification, and reaeration of the
bottom layer (the hypoliminion) is inhibited if not eliminated. The thermocline (the layer
of steep thermal gradient between the epiliminion and hypoliminion) limits not only mix-
ing but also photosynthetic activity as well. The hypolimnion has a base oxygen demand
and benthic matter and the settling of particulate matter, from the epiliminion only adds
to this demand. Therefore, while the demands of DO in the hypoliminion increase during
the period of stratification, inhibition of mixing between the epiliminion and the
hypolimnion and the lack of photosynthetic activity deplete the DO concentrations in the
Environmental Hydraulics 5.7
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.8 Chapter Five
hypolimnion. Finally, a rule of thumb suggests that when water temperature is the pre-
dominant cause of differences in water density a temperature gradient of at least 1

C/m is
required to define the thermocline (Mills et al., 1982).
The density of water can be estimated by

T

s
(5.17)
where water density (kg/m
3
),
T
water density as a function of temperature, and

s
increments in density caused by solids.
5.3.2 Water Density as a Function of Temperature
A number of formulations have been proposed to estimate
T
and among these are

T
999.8452594 6.793952 10
2
Te
9.095290 10
3
Te
2
1.001685 10
4
Te
3
(5.18)
1.120083 10
6
Te
4
6.536332 10
9
Te
5
where T
e
water temperature in C(Gill, 1982).
5.3.3 Water Density as a Function of Dissolved Solids or Salinity and
Suspended Solids
In most cases, data for dissolved solids are in the form of total dissolved solids
(TDS); however, in some cases, salinity may be specified. The density increment for dis-
solved solids can be estimated by

TDS
C
TDS
(8.221 10
4
3.87 10
6
Te 4.99 10
8
Te
2
) (5.19)
(Ford and Johnson, 1983), where C
TDS
concentration of TDS (g/m
3
or mg/L). If the con-
centration of TDS is specified in terms of salinity (Gill, 1982).

SL
C
SL
(0.824493 4.0899 10
3
Te 7.6438 10
5
Te
2
8.2467 10
7
Te
3
5.3875 10
9
Te
4
)
C
SL
1.5
(5.72466 10
3
1.0277 10
4
Te
1.6546 10
6
Te
2
) 4.8314 10
4
C
SL
2
(5.20)
where C
SL
concentration of salinity (kg/m
3
). The density increment for suspended
solids is

ss
C
ss
j
,
(
1.

S
1
G

\
(
,
10
3
(5.21)
where S
G
specific gravity of the suspended sediment. (Ford and Johnson, 1983).
The total density increment caused by solids is then

s
(
TDS
or
SL
)
SS
(5.22)
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.9
5.4 MIXING AND DISPERSION IN OPEN CHANNELS
Turbulent diffusion (mixing) refers to the random scattering of particles in a flow by tur-
bulent motions, whereas dispersion is the scattering of particles by the combined effects
of shear and transverse turbulent diffusion. Shear is the advection of a fluid at different
velocities at different positions within the flow.
When a tracer is injected into a homogeneous channel flow, the advective transport
process can be viewed as composed of three stages. In the first stage, the tracer is diluted by
the flow in the channel because of its initial momentum. In the second stage, the tracer is
mixed throughout the cross section by turbulent transport processes. In the third stage, lon-
gitudinal dispersion tends to erase longitudinal variations in the tracer concentration. In
some cases, the second stage is eliminated because the tracer discharge has a significant
amount of initial momentum associated with it; however, in many cases, the tracer flow is
small and the momentum associated with it is insignificant. In the latter case, the first trans-
port stage is eliminated. In this treatment, only the second and third transport stages will be
treated, with the implied assumption that if there is a first stage, it can be treated separately.
The reader is cautioned that, in this chapter, y is the vertical coordinate direction and z
is the transverse coordinate direction.
5.4.1 Vertical Turbulent Diffusion
To develop a quantitative expression for the vertical turbulent diffusion coefficient,
consider a relatively shallow flow in a wide rectangular channel. It can be shown that the
vertical transport of momentum in such a flow is given by

v

d
d
v
y

(5.23)
where shear stress at a distance y above the bottom boundary, fluid density,
v

vertical turbulent diffusion coefficient, and v longitudinal velocity (French, 1985).


Because the one-dimensional vertical velocity profile and shear distribution are known, it
can be shown that

v
kv
*
y
d
j
,
(

y
y
d

\
(
,
j
,
(
1

y
y
d

\
(
,
(5.24)
where k von Karmans turbulence constant (0.41), y
d
depth of flow, v
*
shear veloc-
ity ( gy
d
S), and S longitudinal channel slope (French, 1985). The depthaveraged
value of
v
is

v
0.067y
d
v
*
(5.25)
When the fluid is stably stratified, mixing in the vertical direction is inhibited, and one
often quoted formula expressing the relationship between the unstratified and stratified
vertical mixing coefficient was provided by Munk and Anderson (1948):

vs

1 3.3

v
3 Ri)
1.5

(5.26)
where
vs
the stratified vertical mixing coefficient.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.4.2 Transverse Turbulent Diffusion
In the infinitely wide channel hypothesized to derive Eq. (5.24), there is no transverse
velocity profile; therefore, a quantitative expression for
t
, the transverse turbulent diffu-
sion coefficient, cannot be derived from theory. The following equations to estimate
t
derived from experiments by Fischer et al., (1979), and Lau and Krishnappen (1977).
In straight rectangular channels, an approximate average of the results available is

t
0.15y
d
v
*
50% (5.27)
where the 50 percent indicates the error incurred in estimating
t
. In natural channels,
t
is significantly greater than the value estimated by Eq. (5.27). For channels that can be
classified as slowly meandering with only moderate boundary irregularities

t
0.60y
d
v
*
50% (5.28)
If the channel has curves of small radii, rapid changes in channel geometry, or severe
bank irregularities, then the value of
t
will be larger than that estimated by Eq. (5.28).
For example, in the case of meanders, Fischer (1969) estimated that

t
25

V
R
2
2
y
c
v
3
d

*
(5.29)
where a slowly meandering channel is one in which

R
T
c
V
v
*

2 (5.30)
and R
c
radius of the curve.
As stated above, the complete advective transport process in a two-dimensional flow
can be conveniently viewed as composed of three stages. In the second stage, the prima-
ry transport mechanism is turbulent diffusion, and a comparison of Eqs. (5.25) and (5.27)
shows that the rate of transverse mixing is roughly 10 times greater than the rate of verti-
cal mixing. Thus, the rate at which a plume of tracer spreads laterally is an order of mag-
nitude larger than is the rate of spread in the vertical direction. However, most channels
are much wider than they are deep. In a typical case, it will take approximately 90 times
as long for a plume to spread completely across the channel as it will take to mix in the
vertical dimension. Therefore, in most applications, it is appropriate to begin by assuming
that the tracer is uniformly distributed over the vertical.
In a diffusional process in which the tracer is added at a constant mass flow rate (M
*
)
at the center line of a bounded channel (C/z 0 at z 0 and C/z 0 at z T), the
downstream concentration of tracer is given approximately by

C
C

4
1
x'

exp
j
,
(

(z' 2
4
n
x
'
z
o
')
2

\
(
,
exp
j
,
(

(z' 2
4
n
x'
z
o
')
2

\
(
,

(5.31)
where
C'

V
M
T
*
y
d

5.10 Chapter Five


Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
x'

V
x
T

t
2

and
z'

T
z

A reasonable criterion for the distance required for complete mixing (where the con-
centration is within 5 percent of its mean value everywhere in the cross section) from a
center-line discharge is
L

0.1

V
t
T
2

(5.32)
If the pollutant is discharged at the side of the channel, the width over which the mix-
ing must take place is twice that for center-line injection, but the boundary conditions are
otherwise identical and Eq. (5.32) applies if T is replaced with 2T.
5.4.3 Longitudinal Dispersion
After a tracer becomes mixed across the cross section, the final stage in the mixing process
is the reduction of longitudinal gradients by dispersion. If a conservative tracer is dis-
charged at a constant rate into a channel, the flow rate of which also is constant, there is
no need to be concerned about dispersion; however, in the case of an accidental release
(spill) of a tracer into a channel or the release is cyclic, dispersion is important. The one-
dimensional equation governing longitudinal dispersion is

C
t
V

C
x
K

2
2
C
x
S (5.33)
where K the longitudinal dispersion coefficient and S sources or sinks of materials.
The initial work in dispersion, beginning with Taylor (1954), assumed a prismatic chan-
nel. However, natural streams have bends, sandbars, side pools, in-channel pools, bridge
piers, and other natural and anthropogenic changes, and every irregularity in the channel
contributes to longitudinal dispersion. Some channels may be so irregular that no reason-
able approximation of dispersion is possible: for example, a mountain stream consisting
of pools and riffles.
Fischer et al. (1979) presented a number of methods of approximating K in a natural
open channel. Of these, the most practical is
K

0.01
y
1
d
v
V
*
2
T
2

(5.34)
Equation (5.33) depends on a crude estimate of
t
and does not reflect the existence of
dead zones in natural channels. However, it does have the advantage of relying only on
the usually available estimates of depth, velocity, width, and surface slope.
With regard to the solution of the dispersion equation, the following observations are
pertinent:
1. The longitudinal dispersion analysis is not valid until the end of the initial period,
when
x

0.4

V
t
T
2

(5.35)
Environmental Hydraulics 5.11
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.12 Chapter Five
2. In the case of a slug of dispersing material (mass M), the longitudinal length of the
cloud after the initial period can be estimated approximately by
L 4
,
,

2K

t
T
2

j
,
(

V
x
T

t
2

0.07
\
(
,
]
]
]
0.5
(5.36)
and the peak concentration within the dispersing cloud is
C
max
(5.37)
Note that the observed value of the peak concentration will generally be less than this
estimate because some of the material is trapped in dead zones and some of the typical
tracers (Martin and Mc Cutcheon, 1999) sorb onto sediment particles.
5.5 MIXING DISPERSION IN LAKES AND RESERVOIRS
Important factors in the hydraulic design, operation, and analysis of spills in reservoirs
and lakes include (1) determining vertical stratification to guide lake monitoring and the
design withdrawal structures, (2) locating the plunge point or separation point to deter-
mine how inflows mix, (3) computing the dilution and mixing of inflows and the time
required to travel through a reservoir or lake, and (4) determining the quality of with-
drawals or outflows and effects on the quality of reservoir water. The elevation and flow
through withdrawal structures at dams are selected to control flooding and achieve cer-
tain water-quality targets or standards. The stratification, mixing, and travel of inflows
are determined to design water-intake structures at dams or other locations in lakes, to
forecast the habitat and fisheries that a proposed reservoir may support, and to track
chemical spills or flood waters through reservoirs. This section is based on
Chaps. 8 and 9 in Martin and McCutcheon (1999), which provide a number of sample
calculations.
Many lakes and reservoirs stratify for part of the year into an epilimnion, thermocline,
and hypolimnion illustrated in Fig. 5.3. The depth and thickness of the thermocline or met-
alimnion vary with location and time of the year and even time of the day to a limited
extent. The thermocline represents the interface between a well-mixed surface layer, or
epilimnion, and the cooler, deeper hypolimnion. In freshwater lakes, the thermocline is
defined by a minimum temperature gradient of 1C/m. When a distinct interface does not
exist, the thermocline, epilimnion, and hypolimnion may not be defined. Mixing process-
es also are different in riverine, transition, and lacustrine zones (Fig. 5.3). Mixing in the
riverine zone is dominated by advection and bottom shear, and turbulence is generally dis-
sipated under the same conditions. Seiche, wind mixing, boundary shear, boundary intru-
sion, withdrawal shear, internal waves, and dissipation of turbulence generated elsewhere
cause mixing in the lacustrine zone. Buoyancy resulting from stable stratification stabi-
lizes or prevents mixing. In the transition zone, ending at the plunge point or separation
point, buoyancy begins to balance the advective force of the inflow. There are three
sources of energy for mixing: (1) inflows from tributaries, overland runoff, and dis-
charges, (2) withdrawal at dams, discharges at control structures, and natural outflows,
and (3) wind shear, solar heating and cooling, heat conduction and evaporation, and other
meteorological forces.
M

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.13
F
I
G
U
R
E

5
.
3
M
i
x
i
n
g

p
r
o
c
e
s
s
e
s

i
n

z
o
n
e
s

o
f

l
a
k
e
s

a
n
d

r
e
s
e
r
v
o
i
r
s
.

(
M
o
d
i
f
i
e
d

f
r
o
m

F
i
s
c
h
e
r
,
e
t

a
l
.
1
9
7
9
)
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Shallow lakes and reservoirs that do not stratify are normally analyzed in the same
fashion as rivers or as a completely mixed body of water. For a completely mixed system,
the residence time (T in seconds or more typically years) or time for an inflow to travel
through the body of water is simply t
r
/Q, where is the volume of the lake (m
3
) and
Q is the sum of the inflows or the average reservoir discharge (m
3
/s).
Freshwater lakes tend to stratify when the mean depth exceeds 10 m and the residence
time exceeds 20 days (Ford and Johnson, 1986). The densimetric or internal Froude num-
ber Fr
d
(Norton et al., 1968) provides a better indication of the stratification potential of a
reservoir where
Fr
d
Fr
p
(5.38)
L
L
the length of the reservoir (m), y
avg
the its mean depth (m), g gravitational accel-
eration (m/s
2
), the difference in density over the depth for the internal Fr or between
the inflow and surface waters of the lake or reservoir at the plunge point or separation
point (kg /m
3
), r average density of the lake for the internal Fr or density of the inflow
(Turner, 1973) at plunge or separation points (kg/m
3
), V
o
the average velocity of the
inflow (m/s), and y
o
the hydraulic depth or cross-sectional area divided by the top width
of the inflow (m). The Fr at the plunge point Fr
p
, also defined in Eq. (5.38), will be used
in the next section. For design projections, the dimensionless density gradient /(y
avg
)
normally is taken to be 10
-6
m
1
(Norton et al., 1968). If Fr >> 1/, the reservoir is expect-
ed to be well mixed. If Fr << 1/, the reservoir is expected to be strongly stratified, and
when Fr 1/, the reservoir is expected to be weakly or intermittently stratified.
Using the length or depth scales of Sundaram (1973) and Ford and Johnson (1986) the
depth to which wind can mix and destroy the stratification of a lake or reservoir for a par-
ticular surface heat flux is
D
t

C
g
p
w
H
3
*
n

5.9 10
9

H
w
3
*
n

(5.39)
where w
*
= the shear velocity of the wind (m/s), = an empirical coefficient approxi-
mately equal to the von Karman constant of 0.4, = the volumetric coefficient of thermal
expansion for water (1.8 10
4
/C), H
n
surface heat flux (W/m
2
), = the density of
water ( 1000 kg/m
3
), and C
p
= the specific heat of water (4186J/kgC).
The wind shear velocity is
w
*

a
C

2
w

1.27 10
3
u
w
, (5.40)
where u
w
= the wind speed (m/s),
a
= the density of air (kg/m
3
), and C
d
= the drag coef-
ficient, which usually is taken to be 1.3 10
3
. (See Martin and McCutcheon (1998) to
estimate the net thermal energy flux.)
5.5.1 Annual Stratification Cycle
In freshwater lakes, stratification results when the sun heats the water faster than wind
shear can mix the heat over the depth. In saline lakes, differences in both temperature and
dissolved solids cause stratification. Stratification involving salinity may persist year-
round in deeper saline lakes. The onset of stratification in freshwater lakes occurs in late
V
o

y
o
L
L
Q

y
avg

5.14 Chapter Five


Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
spring or early summer and persists into the fall or early winter, depending on latitude.
The surface heats rapidly, becoming less dense than deeper layers and forming stable dif-
ferences in vertical density that inhibit vertical mixing until the fall overturn. As stratifi-
cation develops, wind and currents mix the upper layers and tend to deepen the thermo-
cline to form the well-mixed epilimnion. Although storms in late spring and summer
episodically lower the thermocline, the thermocline generally rises as solar heating
increases until midsummer. After later summer cooling begins, the thermocline deepens
until the fall overturn occurs. The decreased difference in temperature in the fall with the
hypolimnion allows more mixing that deepens the epilimnion and thermocline. The vari-
able depth of the thermocline at any time is controlled by seasonal climate, the occurrence
of storms, water temperature, water depth, lake bathymetry, the strength of inflow and out-
flow current, and other factors covered in more detail by Chapra and Reckhow (1983),
Ford and Johnson (1986), Hutchinson (1957), and Wetzel (1983).
The onset of cooler fall conditions causes the epilimnion to lose heat to the atmos-
phere. As heat is lost, mixing tends to become more dominant. The overturning or com-
plete mixing of the reservoir or lake dominates as the epilimnion and hypolimnion
approach the same temperature. During winter, lakes and reservoirs remain unstratified
except in the higher latitudes where the hypolimnion approaches 4C and the surface
approaches 0C. The slight winter stratification of these colder water bodies is the result
of to the usual decrease in water density as temperature decreases from 4 to 0C. Ice cover
maximizes and prevents wind mixing and erosion of the mild differencesm in density.
Stratification is so mild that a distinct thermocline does not form and the epilimnion and
hypolimnion are not well defined. Winter stratification persists until spring warming melts
the ice and heats the surface layer to the temperature of the hypolimnion (usually 4C)
when the spring overturn occurs.
The arrival of spring begins the cycle of heating and stratification anew. A difference
in temperature of just a few degrees results in a difference in density sufficient to inhibit
or prevent most vertical mixing in lakes and reservoirs. Vertical mixing is inhibited almost
completely during summer heating because wind and inflows and outflows do not have
sufficient energy to erode the differences in density that arise. The wind and energy avail-
able from wind and currents cannot overcome the potential energy differences that tend to
prevent mixing of the denser hypolimnion and lighter epilimnion. Fresh water flows into
a saline lake cause salinity gradients that have the same damping effect. Density stratifi-
cation also is caused by suspended sediments, primarily resulting in sediment-laden
underflows. Martin and McCutcheon (1998) have illustrated the stratification cycle for
warmwater lakes and reservoirs.
Run-of-the-river reservoirs and shallow lakes that are weakly stratified because of high
flows or wind mixing, follow only the general stratification trend. Complete mixing may
occur during the summer stratification period as a result of wind or runoff events, and the
thermocline may be difficult to define. Fall overturn occurs earlier in these bodies of water
than it does in deeper lakes.
5.5.2 Plunge and Separation PointEnd of the Transition
Between Riverine and Lacustrine Conditions
The plunge point or separation point marks the downstream end of the transition zone
defined where buoyancy begins to exceed advective forces. These points move seasonal-
ly and, to a limited degree, during the day. Usually distinguished by a line of foam or float-
ing debris across the reservoir or lake, the plunge point occurs when a denser inflow dives
below the lakes surface and continues to flow along the bottom as a density current. The
separation point occurs when an underflow has the same density of the lake water at a
Environmental Hydraulics 5.15
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
given depth, and separates from the bottom to flow into a discrete layer of the lake as an
interflow. Some underflows may be dense enough to flow to the lowest point in a lake or
to the dam that forms a reservoir. If the inflow is less dense than water at the lakes sur-
face, an overflow occurs. Fig. 5.3 illustrates these three types of inflows.
At the plunge or separation point, the internal Fr of the stratified lake Fr
d
is equal to
the Fr of the inflow at that point (Fr
p
), as noted in Eq. Fr
d
(5.38). If the difference in den-
sity in between the lakes surface and the inflow is positive, an overflow occurs, and
if is negative, an underflow occurs. If the slope of the reservoir bottom, or valley, is
mild (S
B
0.007), then the hydraulic depth (y
o
) is the normal depth of flow. For steep
slopes (S
B
0.007), the hydraulic depth is the critical depth (Akiyama and Stefan, 1984).
For tributary or river channels that are approximately rectangular or triangular, the
hydraulic depth and location of the plunge point or separation point can be calculated. For
a rectangular cross section of constant width, the hydraulic depth is
y
o


1/3


1/3
(5.41)
where Q the riverine inflow rate (m
3
/s) equal to VA, A the cross-sectional flow in area
of the river (m
2
), B the conveyance width (m), and q the flow per unit width (m
2
/s).
Similar expressions were proposed by Akiyama and Stefan (1984), Jain (1981), Singh and
Shah (1971), and Wunderlich and Elder (1973), among and others. Savage and Brimberg
(1973) developed an independent expression for the Froude number at the plunge point or
point of separation (Fr
p
) based on the conservation of energy and the theory of two-lay-
ered flow in stratified water bodies, which can be expressed as
Fr
p

j
,
(

S
f
b
b

\
(
,
0.478
(5.42)
where f
b
the dimensionless bed friction factor and f
i
dimensionless interfacial fric-
tion. Martin and McCutcheon (1998) have illustrated the calculations and summarized the
validation of these equations by an example derived from Ford and Johnson (1981, 1983).
For a triangular cross section with an angle 2 between the channel or valley walls, the
hydraulic depth is one-half the total depth. The area of the cross section (m
2
) is A y
2
o
tan(), which, when substituted into the expression for the normal densimetric number Fr
n
and solved for the hydraulic depth y
o
(m), is
y
o
0.5

1\5
where the bottom depth (distance between water surface and apex of the triangular cross
section) is twice the hydraulic depth for a triangular cross section. Hebbert et al. (1979)
derived an expression for the downstream densimetric Froude number at the plunge point
or separation point F
p
for normal flow (S
B
0.007) in a triangular crosssection, related
to the reservoir characteristics as
2Q
2

Fr
n
2

tan
2
()

2.05

f
f
b
i

q
2

Fr
p
2
g

Q
2

Fr
2
p
g

B
2
5.16 Chapter Five
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Fr
2
n

sin()
C
t
D
an(S
b
)

[1 0.85 C
1/2
D
sin ()] (5.44)
where C
D
the dimensionless bottom drag coefficient [C
D
(f
i
f
b
)/4].
Equations (5.42) and (5.44) are based on characteristics of the reservoir or lake. (See
Martin and McCutcheon (1998) and Gu et al. (1996) for an example of the calcula-
tions.)
5.5.3 Speed, Thickness, and Width of Overflows
Martin and McCutcheon (1998) have noted that the speed of an overflow (v
of
with dimen-
sions m/s) can be estimated from the celerity of a wave in a frictionless flow, but this con-
sistently overestimates the rate of spread. Instead, Koh (1976) developed a more practi-
cal semiempirical expression based on uniform flow which reduces to (Ford and
Johnson, 1983)
v
of
1.04

of
where the thickness of the overflow y
of
(m) can be estimated from (Kao, 1976) as
y
of
1.24
,
,

]
]
]
1/3
(5.46)
In natural settings, overflows are usually dissipated by mixing caused by wind and solar
heating before traveling too far.
Horizontal spreading of an overflow is estimated using the inflow Fr defined by Eq.
(5.38). Safaie (1979, cited in Ford and Johnson, 1983) found that for Fr
d
3, the flow is
an unsteady, buoyancy-driven spread and can be assumed to be completely mixed lateral-
ly except for abrupt changes in the entrance geometry. Typically, reservoirs widen gradu-
ally where major tributaries enter, but lakes may have an abrupt widening at the mouth of
tributaries. For Fr
d
3, the inflow acts like a jet that expands proportionally with distance
B(x) B
0
cx where B(x) the overflow width (m) at distance x measured from the sep-
aration point (m), B
0
the width of the riverine or tributary flow at the separation point
(m), and c a dimensionless empirical constant (Ford and Johnson, 1983). From labora-
tory experiments with plane jets, the value of c has been determined to be approximately
0.16 (Fischer et al. 1979; Ford and Johnson, 1983).
5.5.4 Underflow or Density Current Mixing
Underflows are dominated by two mixing processes. First, significant mixing occurs dur-
ing the plunge beneath the surface. Second, shear at the interface with ambient lake or
reservoir water will result in mixing and entrainment as the underflow moves downward.
The initial turbulent mixing of the plunging flow will increase the total flow rate of the
underflow and reduce the density and concentration gradients. The fraction entrainment
caused by plunging is (Q
p
Q)/Q, where Q
p
is the flow rate at the plunge point (m
3
/s) and
Q is the river flowrate (m
3
/s). For mild slopes S
B
< 0.007, is on the order of 0.15
(Akiyama and Stefan, 1984). The depth of the underflow is the normal depth of flow. For
steep slopes S
B
> 0.007, is on the order of 1.18 and the density current depth is the crit-
q
2

Environmental Hydraulics 5.17


Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.18 Chapter Five
ical depth (Akiyama and Stefan, 1984). However, the entrained fraction is highly vari-
able. The dilution of concentrations or temperatures resulting from mixing in plunging
flows follows from a simple mass or heat balance
C
p

C
1
a

C
(5.47)
where C is the inflow concentration (g/m
3
or mg/L
3
) or temperature (C), C
a
is the ambi-
ent concentration (g/m
3
or mg/L
3
) or temperature (C) of the lake, and C
p
is the concen-
tration (g/m
3
or mg/L
3
) or temperature (C) of the plunging flow after initial mixing.
The mixing after plunging results from bottom shear as well as shear at the interface
of the underflow with ambient lake water. For a triangular cross section, the entrainment
coefficient is (Imberger and Patterson, 1981).
E

1
2

C
k
C
3/2
D
Fr
2
b
(5.48)
where laboratory experiments indicate that C
k
is approximately 3.2 (Hebbert et al. 1979),
C
D
the dimensionless bottom drag coefficient defined following Eq (5.42), Fr
b
the inter-
nal fronde number
Fr
b

b
b
h
b

(5.49)
where u
b
underflow velocity, h
b
underflow depth, and
b
relative density differ-
ence. The entrainment coefficient E is a constant for a specific body of water.
y
uf
(6/5)Ex y
O
The depth or thickness of the underflow (m) is a linear function of the entrainment
coefficient (Hebbert et al. 1979; Imberger and, 1981), where x is the distance downstream
from the plunge point (m) and y
o
is the initial thickness of the underflow (m) that is
approximately equal to the depth at the plunge point. If entrainment is limited, the depth
of the underflow remains approximately constant as long as the bottom slope remains con-
stant. The increase in flow rate because of entrainment for an underflow in a triangular
cross section is solved iteratively as
Q(x) Q
1
,
,

j
,
(

y
y
u
1

\
(
,
5/3
1
]
]
] (5.50)
where Q
1
the discharge (m
3
/s) and y
1
the depth (m) from the previous calculation
step. For the initial iteration, Q
1
the discharge at the plunge point Q
p
(m
3
/s) and y
1

the plunge point depth y


o
(m).
Because of more significant differences in density and less internal mixing contrasted
with the epilimnion, underflows tend to remain more coherent than overflows. Sediment
laden underflows, especially, tend to travel to the lake outlet or dam.
5.5.5 Interflow Mixing
After experiencing approximately 15 percent entrainment at the plunge point (for mild
slopes) and mixing as an underflow, an interflow intrudes into a lake at the depth at which
neutral buoyancy is achieved. The turbulence generated by bottom shear is dissipated
quickly, and entrainment into the interflow is dominated by interfacial shear with ambient
lake water above and below the intrusion layer.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.19
When the momentum of inflow is small, an interflow is analogous to a withdrawal
from a dam discussed in Sec. (5.5.6). Interflows are governed chiefly by three conditions
based on the dimensionless number R Fr
i
G
r
1/3
where Fr
i
is the internal Fr defined in Eq.
(5.51) and G
r
is the Grashof number (G
r
), both of which are computed at the depth of
intrusion. The internal Froude Number computed at the intrusion depth is
Fr
i

N
q
L
I
2
I

B
I
Q
N
I
L
2
I

(5.51)
where q
I
the interflow rate per unit width following entrainment at the intrusion point
( m
2
/s), L
I
the length of the reservoir at the level of intrusion (m), Q
I
the interflow rate
(m
2
/s), B
I
the intrusion width (m), and N the buoyancy frequency (s
-1
) expressed as
N

I
y

(5.52)
where
I
density difference between the layers into which the flow is intruding
(kg/m
3
),
I
density of the intrusion (kg/m
3
), and y
I
the thickness of the depth of the
intrusion (m). The dimensionless Grashof number G
r
is the square of the ratio of the dis-
sipation time to the internal wave period or
G
r

2
2
L
v
4
I

(5.53)
where
v
the vertically averaged diffusivity (m
2
/s). Generally, if G
r
1, then an inter-
nal wave field will decay slowly, but if G
r
1 then viscous dissipation damps waves
quickly (Fischer et al. 1979). Imberger and Patterson (1981) also introduced a dimen-
sionless time variable
t
*

G
t
r
N
1/6

where t time(s), which, along with the Prandtl number Pr


v
/
t
, where
t
is vertical-
ly averaged diffusivity of heat (m
2
/s), is used to define three interflow conditions:
1. If R 1, the intrusion is governed by a balance of the inertial and buoyancy forces
so that the actual intrusion length L
i
is proportional to time, as given by (Ford and
Johnson, 1983; Imberger et al., 1976).
L
i
0.44L
i
R t
*
0.44 q
I
N t (5.54)
If the speed of the intrusion is constant or uniform, the velocity v
I
is L
i
/t, so that
v
I
0.44 q
I
N 0.194
j
,
(

I
I
y
I

\
(
,
1/2
where
m
the density of the intrusion. The difference in density in the computation
of the buoyancy frequency is that occurring over the thickness of the intrusion h
m
,
which, along with the relationship u
m
q
m
/h
m
, can be substituted into the above equa-
tion to yield an alternative formulation for the speed of intrusion. The thickness of the
interflow can be solved by assuming uniform flow (Ford and Johnson, 1983).
h
m
2.99
,
,

]
]
]
1/3
(5.55)
q
2
m

m
m

Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)


Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.20 Chapter Five
where h
m
is generally distributed equally above and below the center line of the
intrusion.
2. If R t
*
R P
2/3
r
, then the flow regime is dominated by the balance between vis-
cous and buoyancy forces and the intrusion length becomes
L
i
0.57 L R
2/3
t
*
5/6
(5.56)
The thickness of the interflow is
h
m
5.5L
m
G
r
1/6
(5.57)
In this regime, the flow is generally distributed so that 64 percent lies above the cen-
ter line of the intrusion (Imberger, 1980); thus, the half-thickness (h
ma
) of the interflow
above the center line is given by
h
ma
3.5L
m
G
r
1/6
(5.58)
and the half-thickness below the center line is given by
h
mb
2.0L
m
G
r
1/6
(5.59)
3. If P
2/3
r
t
*
R
-1
then the flow regime is dominated by viscosity and diffusion and
the intrusion length becomes
L
i
= C

LR
3/4
t
*
3/4
(5.60)
where C

is a coefficient, that generally is unknown (Fischer et al. 1979).


Ford and Johnson (1986) indicated that unless dissolved solids dominate the densi-
ty profile (i.e., P
r
is high), intrusions into most reservoirs have R 1, where inertia and
buoyancy dominate. Because the difference in density varies with the location of the
limits of the interflow zone above and below its center line, the solution proceeds by
estimating a value of h
m
and then by computing the difference in density, which is then
used to compute a revised estimate of h
m
. This process is repeated until convergence
occurs.
The equations for intrusion require information on both the morphometry of the reser-
voir and the temperature distribution. The widths used in the formulations should repre-
sent the conveyance width (Ford and Johnson, 1983). Because the time for the intrusion
to pass through a lake can be relatively long, the flow rates used in the calculations
should represent an average value over the period of intrusion. To estimate the time scale
in their analysis of intrusions in DeGray Lake in Arkansas, Ford and Johnson (1983) used
the length of the lake and
m
and h
m
across the thermocline. For DeGray Lake, the intru-
sion time scale ranged from 4 to 6 days. Changes in outflow during the period of the
intrusion also can affect the movement through the lake. Interflows may stall and col-
lapse if the inflow or outflow ends. Interflows also may be diverted or mixed because of
changes in meteorological conditions that influence epilimnion mixing and thermocline
depth.
The temperature or density of the interflow will remain constant. However, the inter-
flow will spread laterally and the thickness will increase caused by entrainment of ambi-
ent water. The resulting concentrations can be computed from a mass balance v
in
CBh
n

constant, where v
in
is the velocity of the interflow; C the concentration or temperature;
B the reservoir width, which may vary with distance from the separation or detachment
point; and h
n
the thickness of the interflow.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.21
F
I
G
U
R
E

5
.
4
R
e
s
e
r
v
o
i
r

w
i
t
h
d
r
a
w
a
l
.

(
A
d
a
p
t
e
d

f
r
o
m

M
a
r
t
i
n

a
n
d

M
c
C
u
t
c
h
e
o
n
,
1
9
9
8
)
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.22 Chapter Five
5.5.6 Outflow Mixing
The withdrawal velocity profile is used in models CE-QUAL-R1 (Environmental
Laboratory, 1985) and CE-QUAL-W2 (Cole and Buchak, 1993) and in calculations to pre-
dict the effects of withdrawals on reservoir and tail race water quality. The extent of a
withdrawal zone (Fig. 5.4) strongly depends on the ambient lake stratification and release
rate, location of the withdrawal, and reservoir bathymetry. For a given outflow rate and
location, the withdrawal zone thins as the density gradient increases. Depending on the
degree of stratification, withdrawal rate and location, and other factors related to the
design of the dam and the bathymetry of the reservoir, the withdrawal zone may be thin
or may extend to the reservoir bottom or water surface. Within the withdrawal zone, the
velocity distribution will vary from a maximum velocity to zero at the limits of the zone,
depending on the shape of the density profile. The maximum velocity is not necessarily
centered on the withdrawal port.
A number of methods predict the extent of withdrawal zones and the resulting veloci-
ty distributions. Fischer et al. (1979) described methods of computing withdrawal patterns
similar to those used in the analysis of interflows in the previous section. The Box
Exchange Transport, Temperature, and Ecology of Reservois (BETTER) model and the
SELECT model based on the original work of Bohan and Grace (1973) are the more prac-
tical approaches. The BETTER model, applied to a number of Tennessee Valley Authority
reservoirs, computes the thickness of the withdrawal zone above and below the outlet ele-
vation from y c
w
Q
out
, where Q
out
the total outflow rate and c
w
is a thickness coeffi-
cient. The model assumes a triangular or Gaussian flow distribution to distribute flows
within the withdrawal zone (Bender et al. 1990).
The SELECT model (Davis et al. 1985) computes the in-pool vertical distribution of
outflow and concentrations of water quality constituents, the outlet configuration and
depth, and the discharge rate (Stefan et al. 1989). The SELECT code also is applied as
subroutines in generalized reservoir models, such as CE-QUAL-R1 (Environmental
Laboratory, 1985). The model is based on the following equations.
The theoretical limits of withdrawal (Bohan and Grace, 1973) were modified by Smith
et al. (1985) to include the withdrawal angle as

Z
Q
3
o
N
ut

(5.61)
where Z distance from the port center line to the upper or lower withdrawal limit;
the withdrawal angle (radians); and N the buoyancy frequency [g/(Z)]
1/2
, in which
the difference in density between that at the upper or lower withdrawal limit and at
the port centerline; and the density (kg/m
3
) at the port center line. The convention is
that is positive for stably stratified flows such that (upper limit) (with-
drawal port) or (withdrawal port) (lower limit). The elevation of the water
surface, the bottom, of the reservoir, and the withdrawal port and the density profile must
be known. The equation must be solved iteratively since both the distance from the port
center line Z and the density as a function of Z are unknown. A typical solution procedure
where the upper and lower withdrawal zones can form freely within the reservoir without
interference at the surface or bottom is as follows:
1. Rearrange the equation as Q
out
Z
3
N/ 0.
2. Check to see if interference exists by, first, using Z equal to the distance from the
port,s center line to the surface. Estimate the density at the center line of the with-
drawal port and the water surface and substitute the values into the rearranged
equation. If the solution is not-zero and is positive, surface interference exists.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.23
F
I
G
U
R
E

5
.
5
D
e
f
i
n
i
t
i
o
n

o
f

w
i
t
h
d
r
a
w
a
l

c
h
a
r
a
c
t
e
r
i
s
t
i
c
s
.

(
F
r
o
m

M
a
r
t
i
n

a
n
d

M
c
C
u
t
c
h
e
o
n
,
1
9
9
8
)
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Similarly, substitute the distance from the port center line to the bottom, along with
the density at the bottom of the reservoir, and determine if a bottom interference
exists.
3. If both of the evaluations from Step 2 are negative, the withdrawal zone forms
freely in the reservoir. The limit of the surface withdrawal zone above the port
can be determined by using iterative estimates of values for Z and the density at
the height above the center line until the equation approaches zero to within some
tolerance. The lower limit of withdrawal below the port center line can be deter-
mined in a similar manner.
4. If surface or bottom interference exists, a theoretical withdrawal limit can be
determined using values of Z computed using elevations above the waters sur-
face for surface interference or below the reservoirs bottom for bottom interfer-
ence. However, this solution requires an estimate of density for regions outside
the limits of the reservoir. Davis et al. (1985) estimated these densities by linear
interpolation using the density at the port center line and the density at the sur-
face or bottom of the reservoir.
For the case where one withdrawal limit intersects a boundary and the other does not,
the freely forming withdrawal limit cannot be estimated precisely using the rearranged
equation. Smith et al. (1985) proposed an extension to estimate the limit of the freely
forming layer similar to that described above

Q
N
out

0.125(D

d)
3

,
,

sin
j
,
(

D
D
d

\
(
,

D
D
d

]
]
]
, (5.62)
where d the distance from the port center line to the boundary of interference (m) and
D the distance between the free withdrawal limit and the boundary of interference (m)
shown in Fig. 5.5. The length scale in the buoyancy frequency N is D in place of Z, and
is the difference in the density between that at the surface for withdrawals that extend
to the surface and between the lower free limit or density at the bottom for withdrawals
that extend to the bottom and upper free limit. For consistency with the definition of sta-
ble stratification as positive, the convention is that (surface layer) (free limit)
or (upper free limit) (bottom layer).
Once the limits of withdrawal are established, the distribution of withdrawal veloci-
ty is estimated by dividing the reservoir into layers, the density of which is determined
at the center line of each layer. The computation of the vertical velocity distribution is
based on the location of the maximum velocity, which can be estimated from (Bohan
and Grace, 1973).
Y
L
Hsin
j
,
(
1.57

Z
H
L

\
(
,
2
(5.63)
where Y
L
the distance from the lower limit to the elevation of maximum velocity (m)
shown in Fig. 5.4, H the vertical distance between the upper and lower withdrawal lim-
its (m), and Z
L
the vertical distance between the outlet center line and the lower with-
drawal limit (m). If the withdrawal intersects a physical boundary, the theoretical with-
drawal limit is used, which may be above the waters surface or below the reservoirs bot-
tom. Once the location of the maximum velocity V
max
(m/s) is determined, the normalized
velocity V
N
(I) V(I)/V
max
in each layer I is estimated for withdrawal zones that intersect
a boundary as (Bohan and Grace, 1973).
5.24 Chapter Five
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
V
N
(I) 1
j
,
(

y
Y
(
L
I)

m
(
a
I
x
)

\
(
,
2
(5.64)
or for a withdrawal that does not intersect a boundary
V
N
(I)
j
,
(
1
\
(
,
2
(5.65)
where V(I) the velocity in layer I (m/s), y(I) the vertical distance from the elevation
of maximum velocity to the center line of layer I (m), Y
L
the vertical distance from the
elevation of maximum velocity to the upper or lower withdrawal limit (m) determined by
whether the centerline of layer I is above or below the point of maximum velocity, (I)
the density difference between the elevation of maximum velocity and the center line
of layer I, and
max
the difference in density between the point of maximum velocity
and the upper or lower withdrawal limit.
If the withdrawal intersects the surface or the bottom, velocities are calculated for loca-
tions either above the waters surface or below the reservoirs bottom and the distribution
is truncated at the reservoirs boundaries to produce the final velocity distribution. The
flow rate in each layer I is
q(I) Q
out
(5.66)
where Q
out
the total release rate and mthe number of layers. The quality of the release
can be determined from a simple flow-weighted average or mass balance as
C
R
V
(5.67)
where C
R
the concentration or temperature of waterquality constituent C in the
release and C(I) the concentration or temperature in each layer.
For discharge over a weir, the withdrawal limit Z and average velocity in the with-
drawal zone V
weir
is derived from the densimetric Froude number [Eq. 5.38] as (Grace,
1971, Martin and McCutcheon, 1998),
0 V
weir

C
1
g
H
(Z
w

H
w
)
2

C
2
(Z

H
w
)

(5.68)
where the difference in density between the weir crest and the lower withdrawal
limit, the density at the weir crest elevation, H
w
head above the weir crest eleva-
tion, Z distance between the crest elevation and the lower withdrawal limit, and C
1
and
C
2
are constants, which have values of
C
1
0.54 and C
2
0 for

Z
H
w
H
w

2.0
and (5.69)
C
1
0.78 and C
2
0.70 for

Z
H
w
H
w

2.0
q(I) C(I)

N
I 1
q(I)
V
N
(I)

m
I = 1
V
N
(I)
y(I) (I)

Y
L

MAX
Environmental Hydraulics 5.25
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.26 Chapter Five
5.5.7 Mixing Caused by Meteorological Forces
Windgenerated waves and convective cooling cause significant mixing at the water sur-
face. Wind shear causes waves at the surface and at each density interface within a lake or
reservoir, such as the thermocline, and larger scale surface mixing by Langmuir circula-
tion results from sustained wind. Wind setup, seiche, and upwelling are caused by mete-
orological events that generate mixing over much larger areas. Internal waves are caused
by shearing currents set up by both wind and other currents and, although not as obvious
as surface waves, these can be larger and more effective in causing mixing. The intensity
of wave mixing and turbulence is a direct result of wind energy or the energy in other
shearing currents.
The basic characteristics of waves are amplitude or height between trough and crest
and the length between crests. The wave period is the time required for successive waves
to pass a given point. Progressive waves move with respect to a fixed point, whereas stand-
ing waves remain stationary while water and air currents move past. The height and peri-
od of wind waves are related to wind speed, duration, and fetch. Fetch is the distance over
which the wind blows or causes shear over the waters surface. As fetch increases, the
wavelength increases; long wavelengths are only produced in the presence of a long fetch.
The shortest wavelengths require only limited contact between wind and water. Waves
with a wavelength less than 2 cm (6.28 cm) are capillary waves, which are not important
in the modeling of lakes and reservoirs. The more important gravity waves have wave-
lengths longer than 2 cm. The two types of gravity waves are short waves and long
waves, distinguished by the interaction with the benthic boundary. The wavelength of
short waves seen by eye on lakes and reservoirs is much less than the waters depth, and
they are not affected by bottom shear. Long waves, such as lake seiche, are influenced by
bottom friction. Seiches are periodic oscillations of the waters surface and density inter-
faces resulting from a displacement.
Shortwave motion is circular in a vertical plane, making a complete revolution as each
successive wave passes. The orbital motion mixes surface layers or layers at an interface.
With no net advection of water, the overall effect is dispersive. Thus, the mixing terms in
transport and water quality models are generally increased to account for wave mixing,
especially in the epilimnion. In a few cases, specific mixinglength formulas Kent and
Pritchard, 1957, Rossby and Montgomery, 1935; were derived for wave mixing, but these
formulas have not been applied in current models of water quality. No appreciable orbital
motion occurs below a depth of approximately one-half the wavelength in unstratified
flow, a depth referred to as the wind mixed depth. The wind mixed depth increases with
fetch because the wave height and wavelength increase with increasing fetch. This is illus-
trated by a simple relationship discovered by Lerman (1978) relating fetch to the depth of
the summer thermocline for a wide variety of lakes of different sizes and shapes.
As wavelength becomes longer in relation to the depth, or as water becomes shallow-
er, wave orbits become increasingly flatter or elliptical. As the orbits flatten, the motion
of the water essentially becomes horizontal oscillation (Smith, 1975) so that the motion
of the water caused by, long waves is more advective rather than dispersive. For long
waves, the wave speed or celerity is c (gY)
0.5
.
As short waves enter shallow water, the bottom affects orbital motion. From this point
inland to the line where wave breaking occurs, the depth is less than onehalf the wave
period. In this shore zone, wave velocity decreases with the square root of the depth,
which results in a corresponding increase in wave height. Waves distort as water at the
crest moves faster than the wave, creating an instability. These unstable waves may even-
tually collapse, forming breakers or whitecaps, depending on the wave steepness of the
waves, the wind speed and direction, the direction of the waves, and the shape and rough-
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.27
ness of the bottom. A spilling breaker tends to form over a gradually shoaling bottom and
tends to break over long distances, with the wave collapsing downward in front of the
wave. Plunging breakers occur when the bottom shoals rapidly or when the direction of
the wind opposes the wave. The plunging breaker begins to curl and then collapses before
the curl is complete. A plunging or surging breaker does actually not break or collapse but
forms a steep peak as the wave moves up the beach. The type of breaking wave and the
associated energy controls beach erosion, aquatic plant growth, surf-zone mixing, and the
exchange of contaminants between surface and ground waters.
After breaking, waves continue to move up a gradually sloping beach until the force of
gravity forces the water back. The extent to which the water runs up the beach is called
the swash zone. The movement of the swash up the beach may result in the deposition of
particles and debris, causing swash marks at the highest point of the zone. Wave runup
in the swash zone also sets up an imbalance of momentum along the porous beach face
that pumps contaminants into and out of the beach (McCutcheon, 1989).
In large lakes and reservoirs with an extremely long fetch, parallel pairs of large verti-
cal vortices or circulatory cells known as Langmuir circulation develop at an angle of 15
clockwise with the general direction of a sustained wind, when wave and current condi-
tions are favorable. The depth of the vortices depends on stratification and may interact
with internal waves formed on the thermocline, deepening over the troughs of internal
waves. Where the counterrotating Langmuir cells converge, visible streaks or bands form
on the surface that tend to accumulate floating debris. In the convergence zone, downward
velocities of 26 cm/s carry surface waters toward the thermocline. These downward cur-
rents move in a circular fashion and turn upward into a divergence zone midway between
the Langmuirstreaks. Water near the thermocline moves to a zone near the surface at a
velocity of about 1 to 2 cm/s over a larger area. As first proposed by Langmuir (1938), this
type of large-scale circulation also contributes to the vertical mixing of the epilimnion.
Like smaller-scale orbital wave mixing, the effect of Langmuir circulation is lumped into
values selected for the eddy viscosities and eddy diffusivities of the epilimnion.
Because of the smaller differences in density across density interfaces within a body of
water, internal waves travel more slowly than do surface waves, but they achieve greater
wave heights. Internal waves include standing waves, such as seiches (Mortimer, 1974) and
internal hydraulic jumps (French, 1985), but most are progressive waves that radiate energy
from the point at which the waves were generated (Ford and Johnson, 1986). Wind shear,
water withdrawals, hydropower releases, and thermal discharges as well as local distur-
bances produce internal waves. The most significant mixing between stratified layers occurs
when internal waves break (Turner, 1973). Before breaking, internal waves mix the water
adjacent to the interface and sharpen the density interface to increase the likelihood of break-
ing. When wave breaking does occur, the entrained water is mixed through the adjacent layer.
Among the most important internal waves is the seiche. As defined above, seiches are
periodic oscillations of the water surface and density interfaces resulting from a displace-
ment. Displacements are typically caused by large scale wind events or large withdrawals.
Sustained wind across a lake surface increases the elevation the waters surface at the
downwind boundary of the lake, causing wind setup. As the wind subsides, the waters
surface tilt or displacement results in a sloshing motion, or seiche, of the lake surface and
in thermocline if the lake is stratified. If hydropower operations or reservoir releases
change the net flow toward the dam, the water piles up at the dam and forms a seiche,
often resulting in noticeable differences in thermocline depths between periods of opera-
tion and nonoperation, such as between weekdays and weekends. More rarely, a seiche
may result from earthquakes or other geologic events. During the rocking or sloshing,
potential energy is converted to kinetic energy and is dissipated by bottom friction.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.28 Chapter Five
Wind setup in Lake Erie may exceed 2 m during severe storms (Wetzel, 1975), but for
a moderate storm blowing over the long axis of Green Bay, Wisconsin the wind setup has
reached approximately 12 cm (Martin and McCutcheon, 1998). An estimate of wind setup
can be obtained from the onedimensional equation of motion assuming constant depth,
negligible bottom stress, and steadystate conditions in an unstratified lake, or

a
g
C

D
y
u
2
w

g
v
y
2
*

(5.70)
where the deviation of the waters surface (m), x the horizontal distance (m),
a

the density of air (kg/m


3
), the density of water (kg/m
3
), C
D
the dimensionless drag
coefficient, u
w
the wind speed (m/s), y the water depth (m), and v
*
the friction
velocity in water (m/s) or (
s
/)
0.5
, in which
s
is the surface shear stress (kg/m s
2
).
The term /x is positive in the direction of the wind. The divergence between wind and
shear force is negligible in shallow lakes and reservoirs but not in deep oceans.
5.6 PLUME AND JET HYDRAULICS
Ajet is the discharge of a fluid from an opening into a large body of the same or similar fluid
that is driven by momentum. A plume is a flow that, while resembling a jet, is the result of
an energy source providing the fluid with positive or negative buoyancy rather than momen-
tum relative to its surroundings. Many discharges into the environment are discussed in
terms of negatively or positively buoyant jets, implying that they derive from sources that
provide both momentum and buoyancy. In such cases, the initial flow is driven primarily by
the momentum of the fluid exiting the opening; however, if the exiting fluid is less or more
dense than the surrounding fluid, it is subsequently acted on by buoyancy forces.
Jets and plumes can be classified as either laminar or turbulent, with the difference
between the two being described by a Reynolds number, as with pipe flow. Near the
source of the flow, the flow of a jet or plume is controlled entirely by the primary initial
conditions that include the mean velocity of the jets exit, the geometry of the exit, and the
initial difference in density between the discharge and the surrounding, or ambient, fluid.
Secondary initial conditions include the intensity of the exiting turbulence and the distri-
bution of the velocity. Following Fischer et al. (1979), the factors of prime importance to
jet dynamics can be defined as follows:
1. Mass flux the mass of fluid passing a jet cross section per unit time
mass flux =
a
A
(u)dA (5.71)
where A is the cross-sectional area of the jet and u the time-averaged velocity of the
jet in the axial direction.
2. Momentum flux amount of momentum passing a jet cross section per unit time
momentum flux =
a
A
(u
2
)dA (5.72)
3. Buoyancy flux buoyant or submerged weight of the fluid passing a jet cross sec-
tion per unit of time
buoyancy flux =
a
A
(gu)dA (5.73)
where the difference in density between the surrounding fluid and the fluid in the
jet. It is convenient to define g()/g g' as the effective gravitational acceleration.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.29
5.6.1 Simple Jets
The two dimensional or plane jet issuing from a slot and the round jet issuing from a noz-
zle into a quiescent ambient fluid are among the simplest cases of jets that can be consid-
ered. These jets have been studied extensively, and there is a reasonable understanding of
how they behave. The boundary between the ambient and jet fluids is sharp at any instant,
and if a tracer were present in the jet fluid, time-averaged measurements would show a
Gaussian distribution of tracer concentration (C) across the jet or

C
C
m

exp
,
,

k
j
j
,
(

y
x

\
(
,
2
]
]
]
(5.74)
where the subscript m the value of C on the jet axis, x the distance along the jet axis,
k
j
experimental coefficients, and y the transverse (or radial) distance from the jet
axis. The Gaussian distribution also is valid for the time-averaged velocity profile across
the jet provided that the measurement is taken downstream of the zone of established flow.
In the case of a circular jet, the length of the zone of established flow is approximately 10
orifice diameters downstream.
Downstream of the zone of established flow, the jet continues to expand and the mean
velocity and tracer concentrations decrease. Within the zone of established flow, the
velocity and concentration profiles are self-similar and can be described in terms of a
maximum value (measured at the jets center line) and a measure of the width or, in the
case of the velocity, distribution:

v
v
m

f
j
,
(

b
y
w

\
(
,
(5.75)
where v
m
the value of v on the jets center line, y a coordinate transverse to the jets
axis, and b
w
the value of x at which v is some specified fraction of v
m
(often taken as
either 0.5 or 0.37; Fischer et al. 1979). The functional form of f in Eq, (5.75) is most often
taken as Gaussian.
Almost all the properties of turbulent jets that are important to engineers can be
deduced from dimensional analysis combined with empirical data (Fischer et al. 1979).
These results are summarized in Table 5.2.
5.6.2 Simple Plumes
Because the simple plume has no initial volume or momentum flux (e.g., smoke rising
from a fire), all variables must be a function of only the buoyancy flux (B), the vertical
distance from the origin (y), and the viscosity of the fluid where
B g
j
,
(

\
(
,
Q g
'
o
Q (5.76)
and
o
difference in density between the plume fluid and the ambient fluid and g
o

apparent gravitational acceleration.
Results similar to those for jets are summarized in Table 5.3, and the numerical con-
stants given are from Chen and Rodi (1976).
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
5.30 Chapter Five
TABLE 5.3 Summary of Plume Properties
Parameter Round Plume Plane Plume
Maximum time-averaged (4.7 0.2)B
1/3
y
-1/3
1.66 B
1/3
velocity
v
m
Maximum time-average (9.1 0.5)M B
-1/3
y
-5/3
2.38M B
-1/3
B
-1
tracer concentration
C
m
Volume flux (0.15 0.015)B
1/3
y
5/3
0.34 B
1/3
y
Q
Ratio 1.4 0.2 0.81 0.1
C
m
/C
avg
Source: After Fischer et al. 1979.
TABLE 5.2 Summary of the Properties of Turbulent Jets
Parameter Round Jet Plane Jet
Initial volume flow rate

D
4
2
V
0

b
0
y
0
V
0
Q
o
Initial momentum flux

D
4
2
V
2
0

b
0
y
0
V
2
0
M
o
Characteristic length scale
M
Q

0
0


M
Q
2
0
0

l
Q
Maximum time-averaged
velocity v
m

M
Q

(7.0

0.1)
j
,
(

l
y
Q

\
(
,
v
m

M
Q

(2.41

0.04)
j
,
(

l
y
Q

\
(
,
V
m
Maximum time-averaged
tracer concentration

C
C
m
0

(5.6

0.1)
j
,
(

l
y
Q

\
(
,

C
C
m
0

(2.38

0.04)
j
,
(

l
y
Q

\
(
,
C
m
Mean dilution

Q
Q
0

(0.25

0.01)
j
,
(

l
y
Q

\
(
,
Q\Q
0
(0.50

0.02)
j
,
(

l
y
Q

\
(
,
Ratio
C
m
/C
av
1.4

0.1 1.2

0.1
Source: After Fischer et al. 1979.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
REFERENCES
Akiyama, J. and H. Stefan, Onset of Underflow in Slightly Diverging Channels, Journal of the
Hydraulics Division, American Society of Civil Engineers113(HY7), 1987, pp. 825844.
Akiyama, J., and H. Stefan, Plunging Flow into a Reservoir, Theory, Journal of the Hydraulics
Division, American Society of Civil Enginners, 110(HY4): 484499, 1984.
Anderson, E. R. Energy Budget Studies, Part of Water Loss Investigations Lake Hefner
Studies., U.S. Geological Survey, Professional Paper No. 269, Washington, DC, 1954.
Beeton, A. M., Relationship Between Secchi Disk Readings and Light Penetration in Lake
Huron, American Fisheries Society Transactions, 87:7379, 1958.
Bender, M. D., G. E. Hauser, M. C. Shiao and W. D. Proctor, BETTER: A Two-Dimensional
Reservoir Water Quality Model, Technical Reference Manual and Users Guide, Report No.
WR28-2-590-152, Tennessee Valley Authority, Engineering Laboratory, Norris, TN, 1990.
Bohan, J. P., and J. L. Grace, Jr., Selective Withdrawal from Man-Made Lakes: Hydraulics
Laboratory Investigation, Technical Report No. H-73-4, U.S. Army Waterways Experiment
Station, Vicksburg, MS, 1973.
Bowie, G. L., W. B. Mills, D. B. Porcella, C. L. Campbell, J. R. Pagenkopf, G. L. Rupp, K. L.
Johnson, W. H. Chan, S. A. Gherini, and C. E. Chamberlin, Rates, Constants, and Kinetics
Formulations in Surface Water Quality Modeling, 2nd ed. EPA/600/385/040, U.S.
Environmental Protection Agency, Environmental Research Laboratory, Athens, GA 1985.
Brown, L. C. and T. O. Barnwel, The Enhanced Stream Water Quality Models QUAL2E and
QUAL2E-UNCAS: Documentation and User Manual,. EPA/600/3-87/007, U.S. Environmental
Protection Agency, 1987.
Chapra, S. C., and K. H. Reckhow, Engineering Approaches for Lake Management, Butterworth,
Boston, 1983.
Chen, C. J., and W. Rodi, A Review of Experimental Data of Vertical Turbulent Buoyant Jets,
Hydraulic Research Report No. 193, Iowa Institute of Hydraulics, Iowa City, IA, 1976.
Cole, T. M., and E. M. Buchak, CE-QUAL-W2: A Two Dimensional, Laterally Averaged,
Hydrodynamic and Water Quality Model, Version 2.0, User Manual, Instruction Report No.
EL951, U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS, 1995.
Davis, J. E., J. P. Holland, M. L.Schneider, and S. C. Wilhelms, SELECT, A Numerical, One
Dimensional Model for Selective Withdrawal, Technical Report, U.S. Army Engineer Waterways
Experiment Station, Vicksburg, MS, 1985.
Environmental Laboratory, CE-QUAL-R1: A Numerical One-Dimensional Model of Reservoir
Water Quality, Users Manual, Instruction Report No. E-82-1, U.S. Army Engineer Waterways
Experiment Station, Vicksburg, MS, 1985.
Fischer, H. B., The Effects of Bends on Dispersion in Streams, Water Resources Research, 5
496506, 1969.
Fischer, H. B., E. J. List, R. C. Y. Koh, J. Imberger, and N. H. Brooks, Mixing in Inland and
Coastal Waters, Academic Press, New York, 1979.
Ford, D. E., and M. C. Johnson, An Assessment of Reservoir Mixing Processes, Technical Report
No E-86-7, U.S. Army Engineer Waterways Experiment Station, Vicksburg, MS, 1986.
Ford, D. E., and M. C. Johnson, An Assessment of Reservoir Density Currents and Inflow
Processes, Technical Report No. E-83-7, U.S. Army Engineer Waterways Experiment Station,
Vicksburg, MS, 1983.
Ford, D. E., and M. C. Johnson, Field Observations of Density Currents in Impoundments, in H.
G. Stefan, ed. Proceedings Symposium on Surface Water Impoundments, ASCE, New York, 1981.
French, R. H., Open-Channel Hydraulics, McGraw-Hill, New York, 1985.
French, R. H., J. J. Cooper, and S. Vigg, Secchi Disc Relationships, AWRA, Water Resources
Bulletin, 18 (1):121123, 1982.
Gill, A. E., Appendix 3, Properties of Seawater, in Atmosphere-Ocean Dynamics, Academic
Press, New York, 1982, pp. 599600.
Environmental Hydraulics 5.31
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Gu, R., S. C. McCutcheon, and P. F. Wang, Modeling Reservoir Density Underflow and Interflow
from a Chemical Spill in a River, Water Resources Research, 32:695705, 1996.
Grace, J. L., Jr., "Selective Withdrawal Characteristics of Weirs: Hydraulics Laboratory
Investigation," Technical Report H-71-4, U.S. Army Engineer Waterways Experiment Station,
Vicksburg, MS, 1971.
Hamon, R. W., L. L. Weiss, and W. T. Wilson, Insulation as an Empirical Function of Daily
Sunshine Duration, Monthly Weather Review, 82(6), 1954.
Hebbert, J., J. Imberger, I. Loh, and J. Patterson, Collie River Flow into Wellington Reservoir,,
Journal of the Hydraulics Division, American Society of Civil Engineers 105(HY(5): 533545,
1979.
Hutchinson, G. E., A Treatise on Limnology: Vol. 1. Geography, Physics and Chemistry, John Wiley
Sons, New York, 1957.
Imberger, J., Selective Withdrawal: A Review, in T, Carstens and T. McClimans, eds.,
Proceedings of the 2nd International Symposium on Stratified Flow, International Association for
Hydraulic Research, Tapir, Trondheim, Norway, 1980.
Imberger, J. and J. C. Patterson, Dynamic Reservoir Simulation Model DYRESM:5, in H.B.
Fischer, ed., Transport Models for Inland and Coastal Waters, Academic Press, Orlando, FL,
1981, pp. 310561.
Imberger, J., R. T. Thompson, and C. Fandry, Selective Withdrawal from a Finite Rectangular
Tank, Journal of Fluid Mechanics, 78:489512, 1976.
Jain, S. C., Plunging Phenomena in Reservoirs, in H. G. Stefan, ed., Proceedings of the
Symposium on Surface Water Impoundments, American Society of Civil Engineers, new York,
1981.
Jobson, H. E., Thermal Modeling of Flow in the San Diego Aqueduct, California, and Its Relation
to Evaporation, Professional Paper No. 1122, U.S. Geological Survey, Washington, DC, 1980.
Jobson, H. E., and T. N. Keefer, Modeling Hight Transient Flow, Mass, and Heat Transport in the
Chattahoochee River Near Atlanta, Georgia, Professional Paper No. 1136, U.S. Geological
Survey, Washington, DC, 1979.
Kent, R. E. and D. W. Pritchard, A Test of Mixing Length Theories in a Coastal Plain Estuary,
Journal of Marine Research, 1:456466, 1957.
Kao, T. W., Principal State of Wake Collapse in a Stratified Fluid, Two-Dimensional Theory,
Physics of Fluids, 19:10711074, 1976.
Langmuir, I., Surface Motion of Water Induced by Wind, Science, 87:119123, 1938.
Lau, Y. L., and B. G. Krishnappen, Transverse Dispersion in Rectangular Channels, ASCE,
Journal of the Hydraulics Division, American Society of Civil Engineers 103(HY10):11731189,
1977.
Lerman, A., ed, Lakes: Chemistry, Geology, Physics, Springer-Verlag, New York, 1978.
Martin, J. L. and S. C.McCutcheon, Hydrodynamics and Transport for Water Quality Modeling,
CRC PressBoca Raton, FL, 1999.
McCutcheon, S. C., Water Quality Modeling: Vol. 1. Transport and Surface Exchange in Rivers,
CRC Press, Boca Raton, FL, 1989.
McCutcheon, S. C., Discussion with Harvey Jobson on Windspeed Coefficients for Stream
Temperature Modeling, Memorandum, U.S. Geological Survey, NSTL Station, MS, March 29,
1982.
Mills, W. B., J. D. Dean, D. B. Porcella, S. A. Gherini, R. J. M. Hudson, W. E. Frick, G. L. Rupp,
and G. L. Bowie, Water Quality Assessment: A Screening Procedure for Toxic and Conventional
Pollutants, EPA600/682004b, U.S. Environmental Protection Agency, Athens, GA, 1982.
Mortimer, C. H., "Lake Hydrodynamics," International Association of Applied Limnology
Mitteilungen, 20:124-197, 1974.
Munk, W., and E. R. Anderson, Notes on a Theory of the Thermocline, Journal of Marine
Research, 7:276295, 1948.
5.32 Chapter Five
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS
Environmental Hydraulics 5.33
Norton, W. R., L. A. Roesner, and G. T. Orlob, Mathematical Models for Predicting Thermal
Changes in Impoundments, EPA Water Pollution Control Research Series, U.S. Environmental
Protection Agency, Washington, DC, 1968.
Rossby, G. G. and R. Montgomery, The Layers of Frictional Influence in Wind and Ocean
Currents, Papers Phy. Oc. Meth., III(3), 1935.
Ryan, P. J. and D. R. F. Harleman, An Analytical and Experimental Study of Transient Cooling
Pond Behavior. Technical Report No. 161, R.M. Parsons Laboratory, Massachusetts Dsitributate
of Technology, 1973.
Safaie, B. Mixing of Buoyant Surface Jet over Sloping Bottom, ASCE, Journal Waterway,
Port, Coastal and Ocean Engineering DIvision, 105(WW4): 357373, 1979.
Savage, S. B. and J. Brimberg, Analysis of Plunging Phenomenon of Density Currents in
Reservoirs, IAHR, 00, 13(2): 187204, 1973.
Shanahan, P., Water Temperature Modeling: A Practical Guide, in Proceedings of the U.S.
Environmental Protection Agency Stormwater and Water Quality Users Group Meeting, LAMR,
jounal, Hydraulic, Research, , April 1984.
Shuttleworth, W. J., Evaporation, in D. R. Maidment, ed., Handbook of Hydrology, McGraw-
Hill, New York, 1993.
Singh, B., and C. R. Shah, Plunging Phenomenon of Density Currents in Reservoirs, La Houille
Blanche, 26(1): 59-64, 1971.
Smith, I. R., Turbulence in Lakes and Rivers, Scientific Publication No. 29, Freshwater Biological
Association, Ambleside, Cumbria, UK, 1975.
Smith, J. A., Precipitation, in D. R. Maidment, ed., Handbook of Hydrology, McGraw-Hill, New
York, 1993.
Smith, D. R., S. C. Wilhelms, J. P. Holland, M. S. Dortch, and J. E. Davis, Improved Description of
Selective Withdrawal Through Point Sinks, Technical Report No, E872, U.S. Army Engineer
Waterways Experiment Station, Vicksburg, MS, 1985.
Stefan, H. G., R. B. Ambrose, Jr., and M. S. Dortch, "Formulation of Water Quality Models for
Streams, Lakes and Reservoirs: Modelers Perspective," Miscellaneous Paper E-89-1, .S. Army
Engineer Waterways Experiment Station, Vicksburg, MS., 1989.
Sundaram, T. R., "A Theoretical Model for Seasonal Thermocline Cycle of Deep Temperate
Lakes," Proceedings of the 16th Conference of Great Lakes Research, 1973, pp. 10091025.
Sverdrup, H. U., M. W. Johnson, and R. H. Fleming, The Oceans, PrenticeHall, Englewood Cliffs,
NJ, 1942.
Swinbank, W. C., Longwave Radiation from Clear Skies. Quarterly Journal of the Royal
Meteorological Society, 89: 339-348, 1963.
Taylor, G. I., " The Dispersion of Matter in a Turbulent Flow Through a Pipe," Proceedings of the
Royal Society of London, Series A, 223: 446-468, 1954.
Thackston, E. L., Effect of Geographical Variation on Performance of Recirculating Cooling Ponds,
EPA-660/2-74-085, U.S. Environmental Protection Agency, Corvallis, OR, 1974.
Tennessee Valley Authority,(TVA) Heat and Mass Transfer Between a Water Surface and the
Atmosphere. TN Report No. 14, TVA Water Resources Research Engineering Laboratory,
Norris, TN, 1972.
Turner, J. S., Buoyancy Effects in Fluids, Cambridge University Press, Cambridge, UK, 1973.
Wetzel, R. G., Limnology, Saunders College Publishing, Philadelphia, PA, 1975, and 1983.
Wunderlich, W. O. and R. A. Elder, Mechanics of Flow Through Man-Made Lakes, in Man-Made
Lakes: Their Problems and Environmental Effects, W. C. Ackerman, G. F. White, and E. B.
Worthington, eds., American Geophysical Union, Washington, DC, 1973.
Downloaded from Digital Engineering Library @ McGraw-Hill (www.digitalengineeringlibrary.com)
Copyright 2004 The McGraw-Hill Companies. All rights reserved.
Any use is subject to the Terms of Use as given at the website.
ENVIRONMENTAL HYDRAULICS

Вам также может понравиться