Physical Limnology

Download as pdf or txt
Download as pdf or txt
You are on page 1of 173

ADVANCES IN APPLIED MECHANICS, VOLUME 27

Physical Limnology
JoRG IMBERGER
and
JOHN C. PATTERSON
Centre for Water Research
The University of Western Australia
Nedlands, Western Australia

I. Introduction .......................................................... 303


11. Seasonal Behavior. ................ ................................. 306
111. SurfaceFlwes ........................................................ 32 1
IV. The Surface Layer .................................................... 334
........ 353
epening ................................................. 370
VII. Differential Heating and Cooling ................ 380
VIII. Outflow .............................................................. 391
IX. Inflow ............................................................... 405
X. Mixingbelowthe Surface Layer ........................................ 413
XI. Modeling ............................................................. 422
XII. Reservoir Destratification by Bubble Aerators ..................... 440
XIII. Summary ............................................................. 45 1
Acknowledgements.. .. ................. 455
References ........................................................... 455

1. Introduction

Our understanding of the hydrodynamics of lakes has developed


.rapidly over the last few years. With new, sophisticated instruments, our
powers of observation are greater, giving us deeper insight into the
physical principles governing the motion and mixing of water in a lake.
303
Copyright 0 1990 Academic Press. Inc.
All rights of reproduction in any form reserved.
ISBN 0-12-002027-0
304 Jorg Imberger and John C . Patterson

The hydrodynamics of lakes is intimately connected with the motion of


a stratified fluid affected by external forcing, the instability of these
motions, and the dynamics of turbulence in a stratified environment.
Most lakes become stratified as solar energy is input at the surface.
Through the expansion of the water, this thermal energy is converted to a
mechanical stability, which arrests motion in the lake. The stronger the
thermal insolation, the stronger the stratification and, in general, the
more quiescent the water body. Wind, stream inflow, and the withdrawal
of water all modify the stratification. However, as explored in this
review, there are mechanisms induced by the stratification that cause
active horizontal water transport and that become more intense as
stratification increases.
The dynamical balance within a lake may be likened to an engine; the
disturbing forces work against the potential energy gradient set up by
radiation. The internal dynamics of this engine are extremely inefficient.
A basin scale disturbance must begin large scale basin seiching, which
then degenerates into smaller scale internal waves and intrusions,
ultimately used by a host of instability mechanisms to produce tur-
bulence. This turbulence then induces a buoyancy flux upwards, raising
the overall center of gravity of the lake. Up to 90% of this mechanical
energy is lost to dissipation.
Lake hydrodynamics appears to have attracted research in two main
areas. First, there is a large body of work in the complicated patterns of
seiching in basins of arbitrary shapes, with and without rotation (Hutter,
1984, 1986; Stocker and Hutter 1986). The second focuses more on the
consequences that follow these internal wave motions, particularly stress,
small scale motions, and mixing processes (Fischer et al., 1979; Imberger
and Hamblin, 1982; Imboden et al., 1983; Imberger, 1985; Imberger,
1987).
Interest in small scale mean motions and turbulence as important
components of lake hydrodynamics is recent: reviews of lake hydrodyna-
mics by Mortimer (1974) and Csanady (1982) mention almost nothing
about these aspects, concentrating instead on internal and surface
seiching introduced by wind stresses.
The similarities between physical oceanography and physical limnology
are many, especially in the second category of work, and with this in
mind, the reader is referred to the following excellent reviews: Garrett
and Munk (1979), Gregg and Briscoe (1979), Caldwell (1983), Fritts
(1984), Holloway (1986), Gregg (1987), Hopfinger (1987), Thorpe
Physical Limnology 305

(1987), and Muller and Garwood (1988), which discuss the motion of a
stratified fluid and mixing due to turbulence. There also have been two
recent major conferences dealing with the oceanic internal gravity waves
(Muller and Pujalet, 1984) and the dynamics of the oceanic surface mixed
layer (Muller and Henderson, 1987). Much of this material is relevant to
an understanding of the hydrodynamics of reservoirs and lakes. To avoid
duplication, this review concentrates on smaller scale motions and the
resulting mixing. A description of the action of rotation has therefore
been almost completely omitted since it concerns the dynamics of internal
waves, and it is adequately described in Hutter (1984,1986) and Stocker
and Hutter (1986). Circulation and mixing in ice-covered waters is an
extremely important subject for a large number of lakes, but this has also
been omitted because of the existence of the review by Carmack (1986).
Because of the rapid development of the subject, much of the
description in this review is speculative and concentrates on findings at
the Center for Water Research. Data from the Wellington, Canning, and
Harding Reservoirs in Western Australia are used as illustrations of
processes. These data sets have been used to set the stage for the review
and literature on the dynamics displayed by these data has been
referenced extensively.
The review begins with a discussion of seasonal behavior and the role
of the geographic location of a lake. Through a new, dimensionless
parameter the dynamics of a lake in the tropics at the time of peak
stratification can be compared with the dynamics of a lake in the
temperate zones at the beginning and end of the stratification cycle. This
section is followed by four sections on the dynamics of the surface layer
(formerly called the mixed layer) beginning with a discussion of the
surface fluxes, followed by the vertical dynamics of the surface layer.
Upwelling, differential deepening, and differential heating and cooling
are presented in separate sections as deviations from the vertical picture.
Our knowledge on outflow has increased considerably since the last
review by Imberger (1980) and this material is updated here. There are
also a number of significant developments in our understanding of the
inflow dynamics, particularly where the inflow separates from the river
channel.
Mixing below the surface layer is also discussed. Because of recent
reviews by Gregg (1987), Hopfinger (1987), and Thorpe (1987), in-
dividual mechanisms are not discussed, and the discussion concentrates
on the underlying dynamics contributing towards mixing in the hypolim-
306 Jorg Zmberger and John C. Patterson

nion and the metalimnion, on the classification of turbulent activity, and


on the parametrization of mixing internal to the hypolimnion and
boundary mixing. These sections set the stage for the review in Section
XI of modeling of lakes where stratification is the dominant force. Lastly,
because of eutrophication problems in lakes throughout the world and
the attention they are receiving, we have included a discussion of some
recent advances in the use of bubble aerators in reservoir destratification.
In this review, the words “lake” and “reservoir” are synonomous, with
“lake” defined as an impoundment that stores water from a catchment. A
reservoir is a man-made lake, which differs from a natural lake in some
ways. In a reservoir, the outflow can be controlled, and, because the
throughflow is more rapid, the water residence time is shorter. The
hydrodynamics of the two are not greatly different, unless the through-
flow is so strong that it affects stratification.

II. Seasonal Behavior

A lake or reservoir is a body of water contained in a river valley,


insulated by the bottom but with heat transfers at the surface. Short wave
solar radiation (of wavelength 300nm to 1OOOnm) and long wave
radiation (of wavelength greater than 1OOOnm) emitted by clouds and
atmospheric water vapor heat the water, while evaporation, sensible heat
transfer, and radiation from the surface most often cool the water. The
net heat input from these sources depends on the season and the
meteorological conditions at the time and site; the balance may change
dramatically not only from day to day but from hour to hour.
The net daily heat flux at the surface of the Canning Reservoir, 50 km
southeast of Perth, Western Australia, is shown in Figure 1. The
reservoir is at a latitude of 32.5”s with a Mediterranean climate of cool,
wet winters from June to August, and hot, dry summers from December
to March. The sensible and latent heat fluxes used in Figure 1 are
calculated from meteorological data gathered 2.0 m above the water
surface using bulk aerodynamic flux equations (see Section 111) with
coefficient values applicable to a neutral atmosphere (Tennessee Valley
Authority, 1972; Fischer et al., 1979).
The variance of the daily net heat flux was larger in spring and summer
(September to March) than in autumn and winter (April to August). This
was mainly due to variable cloud cover and higher water surface
Physical Limnology 307
400

200

-
N

2
X

i *
+
z
w

m
YI

-200

-400 I I I 1 I I I I I I I I I I

J J A S O N D J F M A M J J A S

MONTH
1986 1987

FIG. 1. Net heat flux at the surface of the Canning Reservoir. The data were collected
with a meteorological station mounted on a specially designed buoy kept in position with a
taut mooring. The sensors were 2.0 m above the water level at all times. The smooth line
represents the averaged flux, and the tick marks represent the start of the month.

temperature, which led to large back radiation and strong evaporation on


cloudy, windy summer days. Also, the smoothed net heat flux was
negative only for a brief period in autumn (March to May). During
winter, the net flux was approximately zero. In spring and summer, the
smoothed net heat flux was positive with a mean value of approximately
80 W m-*.The total net heat exchange for the year was positive, but the
average lake temperature did not increase between June 1986 and June
1987.
This implies that the sum of the heat flux due to the river inflow, which
tends to be cold, and the loss of heat due to the outflow, which is
somewhat warmer, was comparable in magnitude to the fluxes through
the surface. Inflow and outflow patterns, which differentiate lakes from
reservoirs, must therefore have a significant influence on the heat budget
of a lake (see also Myrup et al. (1979) for a detailed heat budget of Lake
Tahoe).
308 Jorg Imberger and John C. Patterson

50.0

I
J
I
J A
I
S
I
O
I I
N
I
D
I
J
1
F
I
M
1
A
I
M
I
J
I
J A
I

MONM
1986 1987

FIG.2. The thermal characteristicsof the water in the Canning Reservoir corresponding
to the net heat flux shown in Figure 1. Drops were collected at fortnightly intervals
throughout the 15 months.

The Canning Reservoir thermal characteristics corresponding to the


period described in Figure 1, are shown in Figure 2. During winter, cold
weather mixed the lake over the whole depth, leading to an isothermal
water body of approximately 12°C. The inflow during the winter had
much the same temperature, as seen in Figure 2, and the temperature at
the bottom of the lake remained at 12°C. Just before the minimum
temperature was reached at the bottom of the lake, the surface layers
began to develop a weak thermal gradient in direct response to the
positive net heat flux (see Figure 1). This heating intensified until, at the
Physical Limnology 309

end of December, the temperature at the surface had reached 24°C over
a depth of nearly 10 m. The stratification during this period appeared to
be unable to suppress the turbulence in the deep part of the lake between
20 and 30m depth, where the temperature rose from a low just above
11°C in September to 12°C in December. After December, the warmer
surface temperatures increased the back radiation, the evaporation, and
the sensible heat loss, and the surface temperatures rose only slightly,
reaching a maximum of 25°C in late February.
After this time, penetrative convection, introduced by a net surface
cooling, led to a characteristic decrease of the surface water temperature
of the top 15m, which continued from late February to June. Also
noticeable was the gradual deepening of the isotherms at the base of the
surface layer, induced by the erosion at the base by both natural
convection and wind-induced turbulence. Surface cooling has two effects.
First, the kinetic energy of the penetrative convection thermals erodes
the stable structure underlying the surface layer. Second, the cooling
decreases the stability (the density difference between the surface layer
and the underlying water), thus making the base more susceptible to
erosion by turbulence from within the surface layer. By June, the lake
again had become isothermal.
In limnology textbooks (Hutchinson, 1957; Wetzel, 1975; Henderson-
Sellers, 1984b), the upper warm layer is called the epilimnion, the strong
temperature gradient region below is called the metalimnion, and the
colder water in the lowest part of the lake the hypolimnion. These terms
are useful for description, but as lmberger (1985) pointed out, the
understanding of the properties of water behavior in these regions has
recently been revised.
As the following sections will show, the surface layer is not as well
mixed as is generally believed, nor is it in a state of uniform and constant
turbulence. The surface layer, or epilimnion, responds to diurnal changes
in surface fluxes. The metalimnion contains the major temperature
changes between the surface and bottom water. It is made up of many
temperature steps, and it is unrealistic to define the thermocline as that
depth at which the temperature gradient is maximum, as has been
suggested by previous authors (see Wetzel (1975) and Straskraba (1980)).
Typical profiles from the data set of Figure 2 are shown in Figure 3. The
temperature gradient is a strong function of the resolution of the
instrument. For the smoothed profile with a resolution roughly matching
the vertical extent of the interface (Figure 3b), the gradient reaches only
w
b C
5;

L I I I I I I I I 1140
11 18 25 -15 15 -15 0 15 -15 0 15
TEMPERATURE StK dED FINESCALE MICROSTRUCTURE
(.C) dTMz ('C/rn) dT/dz ('C/rn) dT/& ( Q n )

FIG.3. Typical profile data obtained in the Canning Reservoir on 26 January 1987 at a central location. (a) Temperature obtained
with a standard conductivity, temperature, depth profiling instrument. (b) Smoothed temperature gradient. (c) Enhanced temperature
gradient with an effeective resolution of 0.02 m. (d) Temperature gradient obtained with a microstructure profiler which had a resolution
of 1 mm. This profile was taken a little earlier than that shown in (a), (b), and (c) and only extended to a depth of 20m.
Physical Limnology 311

3°C per meter, whereas at the finest resolution obtained with a


temperature microstructure instrument (which resolved the profile to
1mm intervals) a maximum temperature gradient of 15°C per meter was
observed (Figure 3d). The coarse definitions find their origin in the early
measurements of Wedderburn (1906, 1911, 1912) where temperatures
were measured at every one or two meters.
The metalimnion is also poorly defined, for example, given the data in
Figure 3a, it is not clear whether the temperature step at 6 m depth
should be included. For this reason Imberger (1985) and Imberger (1987)
introduced the classification of a diurnal surface layer, a parent thermo-
cline, a metalimnion, and hypolimnion. The parent thermocline is the
temperature step caused by the most severe mixing or deepening event in
the immediate past. The metalimnion is the water body from the parent
thermocline to a depth where the smoothed temperature has reached a
certain percentage of the coldest temperature in the hypolimnion.
The parent thermocline is illustrated in Figure 4 with data collected
during a hot, cloud-free day at the Wellington Reservoir. At 0830 hours,
there is a clear surface layer, uniform in temperature from the surface to
approximately 7.0 m. This layer was mixing (Monismith and Imberger,
1988) by penetrative convection; the parent thermocline is the gradient
region found at 6.5 m. Following this, the weather became extremely hot
(temperature at 1200 hours was 42"C), there was no wind, and the water
surface was glassy all day. By 1500 hours (Figure 4), considerable surface
heating had formed a diurnal thermocline to a depth of approximately
1m, with a temperature rise in excess of 4°C. By 1700 hours a brisk wind
of 6ms-' caused the surface to mix (Figure 4), causing the diurnal
thermocline to deepen toward the parent thermocline. By definition,
once the diurnal thermocline has reached its maximum depth, it becomes
the new parent thermocline. In some cases, this may be deeper than the
previous parent thermocline (see Imberger (1987), Schertzer et al.
(1987), and Strub (1983)). Generally during spring, lakes stratify and the
location of the parent thermocline becomes shallower with time; in
autumn, during the lake's cooling phase, the parent thermocline lies
progressively deeper.
The temperature distribution of a lake depends on the cumulative
effect of the surface fluxes, the inflows, the outflows, and the rate of
vertical mixing. There has, however, been a great effort made (see
Straskraba (1980)) to provide empirical formulae to predict the thermal
characteristics of a lake according to its location. Three parameters are
312 Jorg Zmberger and John C . Patterson

TEMPERATURE (“C)

RG. 4. Evolution of a diurnal surface layer. Data taken in the Wellington Reservoir on
25 February 1985. Solid line: data collected at 0824 hours during calm conditions. Dashed
line: data collected at 1501 hours at time of maximum surface temperature gradient. Dotted
line: data collected at 1701 hours; cooling had commenced.

usually characterized: the surface temperature, the bottom temperature,


and the depth of the seasonal thermocline. Given that the solar incident
radiation is determined, at a particular time, by the latitude of the lake
fp, Straskraba (1980) suggested the following relationship for the surface
temperature:
8, = 28.1 - 0.34 fp’+ (0.54 - 0.045 fp’ + 0.0146 fpt2
- 1.97 x lo4 9’3)sin( z + y ) , (2.1)
where
fp’ = fp - 3.4,

[ 60
y = 240
for fp > 0 (northern hemisphere),
for fp < 0 (southern hemisphere),
and t is the Julian day.
Physical Limnology 313

This formula was derived by curve fitting to the measured surface


temperatures of over 50 lakes, both in the southern and northern
hemispheres. Straskraba (1980) mentions that the expected temperatures
may be higher for shallower lakes and lower for deeper lakes than that
predicted by (2.1). Equation (2.1) was tested for two lakes in Western
Australia: Lake Argyle is very large, with a volume of 5.6 x lo9m3 and a
depth of 45m, located at a latitude of 16.5"s; Wellington Reservoir
(Patterson et al., 1984) has a volume of 186 X 106m2,a crest level depth
of 30m, and is located at 33.5"s. Data from these lakes are shown in
Figure 5. The fit for the Wellington Reservoir is reasonable, but for Lake
Argyle the departure both in the mean annual surface temperature and
the annual variation is severe, so such empirical generalizations must be
treated with caution. However, (2.1) illustrates the general trend with
latitude: tropical lakes are warmer and have smaller annual variability,
while lakes at higher latitudes are colder and have larger annual
variations.

LAKE
ARGYLE >

J F M A M J J A S O N D
MOMH
FIG.5. Mean surface water temperature for the Wellington Reservoir and Lake Argyle.
Solid line: data. Dashed line: empirical relationship (2.1) suggested by Straskraba (1980).
314 Jorg Zmberger and John C . Patterson

The temperature of water at the bottom of a lake is determined either


by the temperature of the coldest inflow (Hebbert et al., 1979) or the
temperature at the time of lake overturn (Robarts et al., 1982).
Straskraba (1980) assumes that the bottom temperature remains ap-
proximately constant, which is true for deep lakes such as Lake Tahoe
(Henderson-Sellers, 1984b), but is not true for lakes with active mixing.
Even deep lakes such as Lake Baldegg (maximum depth 70m) show
definite signs of warming during the summer months (Imboden et al.,
1983).
Ignoring the warming due to mixing, Straskraba (1980) suggested that
the bottom temperatures are given by
28.1 - 0.6 @ for 4 5 40",
e,'[4 for @ > 40".
This does not encompass lakes at extreme latitudes such as Lake Erie
(Schertzer et al., 1987), Lake Ontario (Boyce et al., 1983), or Lake
Babine (Farmer and Carmack, 1982) where temperatures fall below 4°C
in winter. The coldest temperatures given by (2.1) should match the
temperatures given by (2.2), since during the coldest period, most lakes
are isothermal. Again, agreement between (2.1) and (2.2) is only
marginal.
Such empirical generalizations ignore many factors that have a
profound influence on the temperature distribution and heat budget. As
mentioned above, inflows and outflows often contribute significantly to
the thermal budget, and this is illustrated in Figure 6 with data from
Patterson et al. (1984). In 1976, water was withdrawn from Wellington
Dam only from an offtake at mid-depth. In 1977, however, most of the
water was withdrawn from a scour valve at the base of the dam wall. The
mid-depth draw policy intensified the metalimnion, which in turn
increased the stability of the lake and thus reduced mixing throughout
(Fischer et al., 1979). By contrast, the bottom withdrawal reduced the
volume of the cold, salty bottom water, which reduced stability and
allowed surface fluxes to mix the lake much earlier in the year.
These variations cannot be incorporated into a simple empirical
formula, but require an estimate of stability of the lake water. Such
stability measures have long been advocated. Hutchinson (1957) intro-
duced the definition for stability:

s, = ( z - z,)A(z)p(z) dz, (2.3)


Physical Limnology 315

40.001

1975 1976 1977 1978

FIG.6. Field measurements of the thermai characteristics of the Wellington Reservoir


from 1975-1978. The metalimnion present in 1976 was absent in 1977 due to the bottom
scour policy instituted in 1977 (Patterson et af., 1984).

where z is the vertical coordinate from the bottom of the lake, A(z) is the
area of the lake at height z , p(z) is the water density at a height z , z, is
the water depth, and zg is the height to the center of volume of the lake,
defined by

Straskraba (1980) suggests that the depth of the center of the meta-
limnion and the rate of turbulent mixing depend on the magnitude of S,.
He also observed that for a particular lake basin undergoing temperature
stratification given by (2.1) and (2.2), S, should be small near the
equator, rise to a maximum at middle latitudes, and again reach zero at
latitudes of approximately 70"; the actual stability being dependent on the
density of the water which is a highly nonlinear function of the
temperature (UNESCO, 1981). Idso (1973) reintroduced this concept of
stability, but multiplied S, by gravity g and divided (2.3) by A(z,),
removing the absolute size of the lake from the definition of the stability.
However, even with these modifications, S, is still dimensional. Further,
the concept of stability expressed by (2.3) is deficient since it only
accounts for the stabilizing influence of the stratification and not the
destabilizing influence of the disturbances of wind, inflows, and outflows.
The problem can be addressed by generalizing the surface layer
stability criterion first introduced by Sverdrup (1949, developed further
by Hurley Octavio et al. (1977), Spigel and Imberger (1980), Thompson
316 Jorg Imberger and John C. Patterson

and Imberger (1980), and formalized by Imberger and Hamblin (1982) as


the Wedderburn number

W = g’h2
- (2.5)
u2,L ’
where g’ is the modified acceleration due to gravity across the base of the
surface layer, h is the thickness of the surface layer, u , is the water
friction velocity due to wind stress, and L is the fetch length. Here
g’ = g Aplp,, where Ap is the density jump across the base of the
surface layer and po is the hypolimnion density. The friction velocity u, is
defined by (3.1) below. This number is dimensionless and represents, as
explained in Imberger (1985) and Imberger (1987), the ratio of the
baroclinic pressure force g’h2 at the point of upwelling and the surface
force u:L imposed by the wind stress; both forces being calculated per
unit width of the lake. This number will be elaborated upon in Section
IV, but a simple generalization is suggested to describe the behavior of
the lake as a whole.
Consider a general lake with an arbitrary stratification p ( z ) being acted
upon by a general wind field with a surface friction velocity u , ( x , y ) ,
where x and y are the horizontal coordinates embedded in the lake’s
surface. As this wind stress is imposed on the surface layer, there will be
a net force acting to overturn the density structure of the water column.
Taking moments about the center of volume located at zg, we obtain for
equilibrium

POUZ, U ( z m - zg) = (zg - zo)MgS, (2.6)

where zo is the center of gravity of the actual water mass with a density
stratification p ( z ) , M is the total mass of water, and /3 is the angle
subtended to the vertical by the center of mass at the center of volume.
This leads to the ratio we shall call the Lake number

(2.7)

Now noting that zo is defined by


Physical Limnology 317

the Lake number may be rewritten as

Following the work of Spigel and Imberger (1980), the angle /3 may be
fixed so that LN is calculated at the point where upwelling commences,
that is when the metalimnion intersects the surface, so that

(2.10)

where zT is the height to the center of the metalimnion, and the fetch is
scaled with A'/*(z,,,), yielding the relationship

(2.11)

Assuming that the wind stress is constant over the surface, then LN
reduces to

(2.12)

where A. is the surface area of the lake A&). For large Lake numbers,
the stratification will be severe and dominate the forces introduced by the
surface wind stress. Under these circumstances, stratification is expected
to be horizontal, with little or no seiching and little turbulent mixing in
the metalimnion or the hypolimnion. Horizontality of the isotherms at
large LN is illustrated by the data set taken at Canning Reservoir on the
18 January 1987, which is depicted in Figure 7. A weak, warm lense of
water resident in the sheltered upstream part of the basin and a very
weak slope of "C m-l towards the dam wall, attributable to a first
mode seiche, are the only departures from this horizontality. As
discussed in Shay and Imberger (1988), the turbulence levels in the
epilimnion and the hypolimnion were extremely low. At the time, the
wind mean shear velocity (averaged over 24 hours) was 0.0033 m s-',
leading to a value of L , = 86. The full implication of LN with respect to
deep mixing will be discussed in Section X.
The interplay between the temperature distribution variation with
latitude and the changing water density due to the elevated temperatures
318 Jorg Zmberger and John C . Patterson

at lower latitudes may be captured effectively by taking the data


displayed in Figure 2, transposing this through the various latitudes via
(2.1) and (2.2), and then calculating L N from (2.12). This was done
by scaling the actual measured surface temperature by the ratio
8,(t, $2)/0s(t,32.5). The bottom temperatures were adjusted to vary
linearly so that at the equator the bottom temperature equalled the
surface temperature, and at 32.5"s the bottom temperature was equal to
that measured in the Canning Reservoir. Here, t is the time of a
particular profile and $2 is the second latitude. The scaled temperature
profiles were then used to compute L N from (2.12), yielding the
two-dimensional variation of LN shown in Figure 8. The diagram shows

TEMPERATURE ("C)
0.0
- 26.0'

10.0

--E
f 20.0
x

30.0

40.0 470 CA65 CA60 CA50 CA45 CAM CA35 CA30 CM5 CA20 CAiC
A j A a A t u
I I
1000.0 2000.0 3000.0 4000.0 5000.0 6000.0

CUMULATIVE DISTANCE

FIG. 7. (a) Longitudinal transect of isotherms in the Canning Reservoir taken on 18


January 1987. (b) Map of Canning Reservoir showing location of stations.
Physical Limnology

9000

8000

7000

POISON GULLY
6000

- 5000
w
VI
+
a
z0
4
K
400C

8
J
a
s
0

300C

2000
CANNING RESERVOIR

1000

0000

0000 1000 2000 3000 4000 5000


LOCAL CO-ORDINATES (m)

FIG. 7. (conld.)
320 Jorg Zmberger and John C. Patterson

FIG.8. An example of the variation of the Lake number as a function of season and
latitude. The surface was constructed by computing the Lake number from data gathered in
the Canning Reservoir, but which was transposed to different latitudes using the empirical
relationship (2.1) suggested by Straskraba (1980). A constant wind speed of 8.0ms-' was
applied throughout.

that the maximum stability is obtained at mid-latitude, around the


summer period. Further, it illustrates that it is possible to trade latitude
and season when analyzing the dynamics of a lake. This would mean that
the Canning Reservoir in August, leaving aside the influence of the
earth's rotation, should be dynamically similar (in the sense of dynamic
similitude) to Lake Argyle in February. More generally, it may be said
that lakes in the tropics behave dynamically similar to lakes at mid-
latitudes at the beginning or end of the stratification cycle.
The seasonal pattern described above is often strongly influenced by
chemicals dissolved in the water, which also influence the density of the
water, thus contributing to the stability of the lake. The origin of these
dissolved salts are threefold (Wetzel, 1975). First, in coastal lakes, sea
water may enter the lake periodically. Second, ground or river salt water
intrusions may be present, and third, salts may be accumulated from the
decomposition of sediments and organic matter. If the dissolved salts
influence the density of water enough to prevent the normal winter
overturn, the lakes are called meromictic. Examples in the literature are
too numerous to list here, and it suffices to mention that they range from
very large lakes, such as the Dead Sea (Steinhorn and Assaf, 1980;
Physical Limnology 321

Steinhorn and Gat, 1983); to small shallow lakes (MacIntyre and Melack,
1982; Bunn and Edward, 1984). The Wellington Reservoir is an example
(see Imberger, 1987) where a strong variability of the salinity of the
inflow led in some years to meromictic behavior.
In general, the dissolved salts influence the dynamics of the lake only
through the equation of state, and provided the Lake number and other
parameters allow for salinity variation, their dynamics are similar to
single species stratification. In the section on deep mixing, we discuss an
exception to this statement, and under certain circumstances, such as in
Lake Kivu (Newman, 1976), double diffusive instabilities may arise in
salt-stratified lakes due to different rates of molecular diffusion between
temperature and salt (see Turner (1985) for a complete review on
multi-components).
Many attempts exist to empirically correlate the dynamics of a lake to
variables such as the surface area (Ward, 1977) and geographical trends
such as altitude (Hutchinson, 1957; Wright, 1961; Loffler, 1968; Khom-
skis, 1969; Lerman and Stiller, 1969; Bella, 1970; Lerman, 1971;
Darbyshire and Colclough, 1972; Khomskis and Filatova, 1972; Blanton,
1973; Idso and Cole, 1973; Lewis, 1973; Sundaram and Rehm, 1973;
Tzur, 1973) and length of the lake (Yoshimura, 1936; Patalas, 1960 and
1961; Arai, 1964; Ventz, 1973). These correlations are useful only in that
they are special cases of the lake number LN formulation.

III. Surface Fluxes

The interaction between the atmosphere and a lake occurs at the lake’s
surface. A great deal of work has been done to determine the fluxes of
momentum, energy and mass across the air-water interface and many
numerical formulae have been developed. These methods are reviewed
by Tennessee Valley Authority (1972), Straskraba (1980), and
Henderson-Sellers (1986).
Short wave radiation (300 nm to 1000 nm) is usually measured directly.
Long wave radiation (greater than 1000nm) emitted from clouds and
atmospheric water vapor can be measured directly or calculated from
cloud cover, air temperature, and humidity (Tennessee Valley Authority,
1972). The reflection coefficient, or albedo, of the short wave and long
wave radiation varies from lake to lake and depends on the angle of the
sun, the color of the water, and the surface wave state. Back radiation
322 Jorg Zmberger and John C. Patterson

from the warm water surface may be calculated from the black body
radiation law, it also may be measured by pointing a radiation instrument
towards the water surface.
It was shown by Strub and Powell (1987), Marti and Imboden (1986),
Keijman (1974), Sadhuram et al., (1988), and many other authors that
simple aerodynamic bulk .,formulae, with constant transfer coefficients,
can be used to calculate the momentum, the sensible heat, and the latent
heat fluxes on a seasonal basis:

E -
-= q ' w ) = -u*q* = -C,U,(q, - q.), (3.3)
PLV
where z is the surface stress, p is the density of the air, u' and w' are the
horizontal and vertical fluctuation of velocity, the overbar is a time
average long enough to average out the short term energy but short
enough to allow for synoptic variability (Busch, 1977; Smith, 1980;
Geernaert et al., 1987), u , is the shear velocity in the air, CD is the
momentum or drag coefficient, U, is the air velocity at a certain height z
above the water surface, H is the sensible heat transfer, Cp is the specific
heat of water, 6" is the temperature fluctuation, 8, is the temperature
scale, CHis the heat transfer coefficient, often called the Stanton number
(Geernaert et al., 1987), 8, is' the water surface temperature, 8, is the
temperature of the air at the height z above the water, E is the latent heat
flux, Lv is the latent heat of vaporization, q' is the specific humidity, q* is
the specific humidity scale, Cw is the latent heat transfer coefficient or
Dalton number (Geernaert et al., 1987), q, is the specific humidity at a
height of z above the water surface, and q. is the specific humidity at
saturation pressure at the water surface temperature.
Equations (3.1), (3.2), and (3.3) are only valid for a stationary surface.
Since the water moves, U, should be replaced by U, - Us, where Us is the
water surface mean velocity (see Businger (1973) for a full discussion),
although for most wind speeds, this is a small correction.
The value of the transfer coefficients CH and Cw for such bulk
modeling has received a great deal of attention and, in the case of lakes,
may be estimated by carrying out a heat and water budget over seasonal
Physical Limnology 323

time scales (for example, Strub and Powell (1987) for Castle Lake, Marti
and Imboden (1986) for Lake Sempach, Myrup el al. (1979) for Lake
Tahoe, and Taylor and Aquise (1984) for Lake Titicaca). These
references suggest a value of CH= Cw = 1.9 X (referenced to 10 m)
yields an adequate description of the total stored energy over a seasonal
time scale. This may be compared with a lower value of 1.45 x
suggested by Hicks (1972). Constant values are sufficient for seasonal
lake modeling since the thermal budget is self-regulating; an underestim-
ate of the heat loss will cause the water surface to heat, thus increasing
heat loss, while an overestimate of heat loss will cause the water surface
to cool and, by (3.2) and (3.3), this will lead to the heat loss being
decreased.
At smaller time scales (from hours to days), this negative feedback
mechanism no longer suffices because the thermal inertia of the water in
the surface layer is too great. The influence of the air column stability
and the water surface roughness must therefore be accounted for, and
these may introduce considerable variability to all three transfer
coefficients. Excellent reviews exist on this topic: Businger (1973), Dyer
(1974), Stewart (1974), Garratt (1977), Wu (1980), Donelan (1982), and
Blanc (1985). There are three main measurement techniques to deter-
mine the instantaneous surface fluxes at the water surface: eddy
correlation, profile, and dissipation. These have been used to calibrate
(3.1) to (3.3) thus giving values for CD, CH, and Cw. A description of
each technique is given by Dobson et al. (1980) and Blanc (1983). The
eddy correlation method is direct; the profile method uses mean velocity,
temperature, and humidity profile data; and the dissipation method uses
the spectral characteristics of the high wave number turbulence. Results
from the eddy correlation technique usually agree with each other within
10% (Blanc 1987); profile techniques usually agree with eddy correlation
methods to within 25% (Miyake et al., 1970; Wucknitz, 1976). Large and
Pond (1981) and Smith and Anderson (1984), in comparisons between
the dissipation method and the eddy correlation technique, found that
the two agreed to within 40%. Within similar schemes, there can be even
greater variability (see Lo and McBean (1978) for interprofile
comparisons).
The reliability of the bulk aerodynamic equations (3.1) to (3.3)
depends on the sensors being located in the internal boundary layer of
the lake. Air coming from land will meet a different roughness, humidity
and temperature over the lake and the internal boundary layer will be
324 Jorg Imberger and John C. Patterson

established with new equilibrium profiles (Taylor, 1970; Venkatram,


1977; Mulhearn, 1981; Claussen, 1987; Garratt, 1987). Peterson (1969)
recommends that the fetch be at least 100 times the height of the sensor,
but Hicks (1975) advocates a more cautious ratio of 1OOO. The recent
simulations of Garratt (1987) suggest that the growth of the internal
thermal internal boundary layer for stable conditions is given by
( 8 )- I y L ,
h = 0 . 0 1 4 x 1 ” Ugmhe
(3.4)

where h is the height of the internal boundary layer, x is the distance


downstream, U, is the free stream air velocity, g is the acceleration due
to gravity, A 8 is the potential temperature difference between the
continental air (constant throughout the approaching boundary layer and
equal to the land surface temperature) and the water surface tempera-
ture, and 8 is the absolute temperature of the air. For a typical
A 8 = 10°C and Um= 5ms-’, this would indicate a layer thickness at
lo00 m of 3.8 m, which lies between the recommendations of Hicks
(1975) and Peterson (1969). Equation (3.4) also agrees well with the
results presented by Mulhearn (1981) and Garratt and Ryan (1988) for
boundary layers close to the shore, even though (3.4) was derived for
considerably larger length scales. The momentum and water vapor
internal boundary layers may be expected to have similar thickness
(Garratt, personal communication).
The unstable case, where the water is warmer than the air, does not
appear to have been solved, but the case of cool ocean air overflowing a
warm land mass was recently reviewed by Venkatram (1986). The
roughness was fixed and not dependent on the shear stress, as in the case
of the model by Garratt (1987), but the internal boundary layer was still
observed to grow as x1I2.
For very small boundary layer thicknesses, the application of equi-
librium profiles severely constrains the position of the instruments, as they
must be placed in that area of the boundary layer described by the law of
the wall (Tennekes and Lumley, 1972) a region that extends only to about
10% of the layer thickness (Sorbjan, 1986). In the example above, where
the layer thickness was 3.8m, the instruments must be placed approxi-
mately 0.4m from the water’s surface. From Donelan (1980), the root
mean square surface roughness due to wave action may be estimated, for
a steady wind, to be 0.02m, which implies a value for Hi13 of O.lm ,
where Hi13 is the significant wave height. The region available for
Physical Limnology 325

positioning of the sensors is thus limited to between 0.1 m and 0.4 m. To


prevent spray damage, the sensors would probably have to be placed
close to 0.4 m from the water surface.
If the meteorological sensors are embedded in this law-of-the-wall
region of the internal boundary layer, and the horizontal advection
effects within the boundary layer can be ignored, the normal turbulence
relationships can be used. Monin and Obukhov (1954) parameterized the
non-dimensional gradients of wind, potential energy, and specific humi-
dity in terms of the stability parameter z l L , where
-pu3,ev
L= (3 * 5 )
LV
+
8, = e ( l 0.61q) is the virtual temperature (“K), k is the von Karman
constant and L is the estimate of the Monin and Obukhov length.
The similarity relationships introduced by Monin and Obukhov (1954)
and Businger (1955), amongst others, are assumed to hold in the
law-of-the-wall region:

kz d e
(3.7)
e, dz
kz dq

Dyer (1974) summarized the various proposed forms for the functions
@ M , @ H , and GW. For a convecting boundary layer ( z / L< 0),

(3.10)

and for a stratified boundary layer with z l L > 0,

Hicks (1976) reanalyzed the Wangara data (Clarke et al., 1971) and
found (3.11) to hold for only weakly stable (0 < z / L < 0.5) boundary
326 Jorg Imberger and John C . Patterson

layers. He suggested that for z / L > 10,


z
&=0.8-, (3.12)
L
whereas Carson and Richards (1978) postulated that for the in-between
region (0.5 < z / L < lo),

@M [
= 8 - 4.25 ):( I'-+):( . (3.13)

Paulson (1970) substituted (3.9) to (3.13) into (3.6) to (3.8) and carried
out the integration to obtain

(3.14)

(3.15)

(3.16)

where zM , zH , and zw are roughness lengths, and defining

= [ - 16(;)] 114
, (3.17)
for the unstable case,
l+x
+ In( y - 2 tan-'(x)
VM= 2 In( 7)
1+x2
) + -,n2 (3.18)

and for the stable case,

-5(3
z
0 < - < 0.5,
L
(3.20)
0.5(i)-2-4.25(;)-1-71n(i)

- 0.852, 0.5 < < 10.0,


L
I
I
(3.21)
z
ln(;) - 0.76(:) - 12.093, - > 10.0.
L
L (3.22)
Physical Limnology 327

These relationships appear to have been tested over the range from
-15 < z / L < 15, but for the values larger than 15, the similarity functions
must be extended as suggested by Sorbjan (1986).
Equations (3.18) to (3.22) can now be used to relate the neutral to the
actual transfer coefficients, leaving only the surface roughness to be
estimated. The roughness lengths are related directly to the drag
coefficients, as can be seen by substituting (3.14) to (3.16) into (3.1) to
(3.4). Under neutral conditions,

(3.23)

(3.24)

(3.25)

For nonneutral conditions, the procedure yields

where a stands for D, H, or W. Geernaert and Katsaros (1986) point out


that the roughness, which is a function of u * , will be influenced by the
stability of the air and the boundary layer so that (3.23) becomes

(3.27)

Now eliminating z and U from (3.1), (3.27), and (3.14) yields

(3.28)

rather than (3.26). The correction (l/k)ln(zMN/zM) is small, seen by


noting that the roughness lengths may be given by the Charnock (1955)
relationship

(3.29)

(3.30)
328 Jorg Zmberger and John C. Patterson

so that (3.28) becomes

(3.31)

The correction at z / L = 1 is about 10%; a reduction for stable conditions


and an increase for unstable conditions. However, Keller et al. (1985)
also investigated the influence of stability on the surface roughness and
found an even greater correction was necessary.
The stability correction can now be implemented in two ways. First,
given the relationships (3.14) to (3.16), these may be substituted into the
definition of the bulk Richardson number (Panofsky, 1963; Deardorff,
1968; Rayner, 1981),
gz A 8 + 0.618,Aq (3.32)
Rig=-(8" U2
to yield a relationship between z / L and Rig that reduces to

(3.33)

where it was assumed that C H N = HWN = CHwN. Deardorff (1968),


however, did not make this assumption, and inverted the equilibrium
form of (3.32) to produce the following relationship (see Strub and
Powell 1987):

(3.34)

(3.35)

C W - e-24 Rie
7 (3.36)
CWN

for stable conditions, and

(3.37)

(3.38)
Physical Limnology 329

for unstable conditions, where

a = 0.83C;:62, (3.39)
b = 0.25C,$'. (3.40)

These relationships were successfully used by Strub and Powell (1987) in


their heat budget calculations at Castle Lake.
The second approach was suggested by Hicks (1975) and involves an
iterative procedure. Neutral values of CDN,CHN and CWNare chosen and
neutral fluxes are calculated by (3.1) to (3.3). These values are
substituted into (3.5) to calculate the Monin and Obukov length L and
thus z / L . New values of the transfer coefficients, now partially corrected
for stability, may be obtained from (3.26). This procedure may be
repeated and the scheme is said to converge when the estimates for L
converge.
This scheme was implemented by Rayner (1981), and some typical
results from the data set presented in Figure 1 are shown in Figures 9a-i.
The air column, judging by the difference in the air and water
temperature shown in Figure 9a was unstable during the nights 29
January and 2 February particularly, but was stable for most of the day
period. The evaporative flux (Figure 9d) is a major contributor to the
heat budget, so the value of Cw is extremely important for an overall
heat budget calculation. The variation of the stability corrected transfer
coefficient for the anemometer height of 2.0m (Figures 9g and h) show
large variations throughout 24-hour periods. The algorithm was bounded
so that -15 < z / L < 15; larger values were not allowed. Dependence on
stability is now generally accepted to be given by (3.26), although direct
verification in terms of CD, CH, and Cw is scarce. For lakes, the reader is
referred to Strub and Powell (1987) for verification under stable
conditions, and to Graf et al. (1984) for both stable and unstable
boundary layers. In general, it appears that the stability corrections
implied by (3.26) describe a diurnal heat flux variation and are necessary
to obtain thermal budget closures in diurnal surface layer modeling
(Imberger 1985).
The remaining question concerns the choice of a neutral drag
coefficient, or as indicated by (3.23) to (3.25), the choice of the surface
roughness z,, zH, and zw. Charnock (1955) put forward, from purely
dimensional grounds, the relationship (3.29) and Stewart (1974) sug-
gested in his review, by fitting data available at the time, a value of (Y
330 b a
33

32

B 31

30

_-11
70
15 19 23 27 31
TEMPERATURE ib)

-Q
2.L

fw/m2)
-
110-31S.W. RADIATION (10-3) NET RADIATION LATENT HEAT FLUX SENSIBLE HEAT FLUX
(W/m2)
(10-3) TOTAL HEAT
FLUX (W/m2)

31 -

30 -

29 0.0 0.8 1.6 2.4 0.0 2.0 4.0 6.0 -15 15


(103)MOMENTUM EXCHANGE (103) HEATNASS TRANSFER STABILITY PARAMETER
COEFFICIENT COEFFICIENT
Physical Limnology 331

between 0.0123 and 0.351. Garratt (1977) added the data from Kitaigor-
odski et al. (1973), Wieringa (1974), Kondo (1975), and Smith and Banke
(1975) to obtain a value of (Y equal to 0.0144; a value that also seemed to
fit inferred drag coefficients applicable to wind speeds up to 50ms-'.
More recently, Wu (1980) arrived at a value of 0.0185. Large and Pond
(1981) derived a value of CDN= 0.012 for wind speeds between 4 and
11m s-l, but increasing with wind speed beyond 11m s-'. Equating the
two formulations at a wind speed of 25 m s-' by using (3.27) leads to a
value of (Y = 0.0097, considerably smaller than suggested by most other
investigators. At low wind speeds (less than 5ms-'), the situation is
unclear, but in general, the results from the above investigators suggest a
constant value of 1 X for C,, (see Hicks (1972)). In general, the
inter-technique variability and data scatter (Blanc, 1985, 1987) outweigh
the above differences, but the most recent values of a; estimated by Wu
(1980), (0.0185); Geernaert et al. (1986), (0.0178); and Geernaert et al.
(1987), (0.0165) appear to suggest a mean value of (Y equal to 0.0175 for
a fully developed wave field.
The large variability in the surface roughness has recently received
considerable attention. Kitaigorodskii (1968) and Kitaigorodskii and
Zaslavskii (1974) already indicated that Charnock's (1955) relationship
(3.29) should be modified to include the wave age Co/u,:

(3.41)

where Co is the phase velocity of the dominant wave (wave number at


maximum energy) and large values of Co/u, represent decaying waves or
large swell and small values indicate a growing sea. This can be rewritten
in terms of the non-dimensional fetch g L / U 2 (Wu, 1985). Many authors
(see Donelan, 1982) have pointed out that for a young sea (short fetch or
small duration), the roughness is greater than under the equivalent
equilibrium conditions and so the neutral drag coefficients may be
expected to be larger. This explains why storm surge modelers (short

FIG. 9. Example of typical heat flux variations computed from data shown in Figure 1.
Day 29 is 29 January 1987. (a) Surface water temperature. (b) Air temperature. (c)
Short-wave incoming radiation. (d) Net total radiation. (e) Latent heat flux. (f) Sensible
heat flux. (g) Total heat flux at the water surface. (h) Drag coefficient C,. (i) Heat and
mass transfer coefficients CH= C, . (j) The bulk Richardson number.
332 Jorg Imberger and John C . Patterson

duration winds) generally require higher values for the neutral drag
coefficient.
Hsu (1974, 1986) and Geernaert et al. (1987) derived a relationship
between the wave age and wave steepness:

(3.42)

where is defined above and Lw is the wave length at the peak of the
energy spectrum. By using both statistics from the JONSWAP formula-
tion and from Chapter 3, Section 4, of the U.S. Army Corps of Engineers
(1977), the wave steepness for equilibrium seas becomes 0.023, yielding a
value of a slightly higher than that proposed by Wu (1980) (ais equal to
0.0185). Graf et al. (1984) investigated drag coefficients for shoaling
waves and large wave steepness and showed that water roughness using
(3.42) yielded good results. Geernaert et al. (1987) carried out a direct
correlation between the neutral drag coefficient C D N and the wave age:

indicating again an increased coefficient for small values of the wave age
(a developing wave surface). The relationships (3.41), (3.42), and (3.43)
also explain the reduction of drag coefficient in shallow water where
Co/u, tends to be larger (Hicks et al., 1974; Emmanuel, 1975; Geernaert
et al., 1987).
Donelan (1982) partitioned the water surface roughness into two parts:

2, = p[ E ( w )dw]lD, (3.44)
and

(3.45)

where wp is the frequency at the peak of the spectrum and /Iis a constant
equal to 0.0125. He then defined a neutral drag coefficient
Physical Limnology 333

where c is the drag coefficient for an immobile plate with roughness


either z, or z, given by (3.27). Here UN is the neutral wind velocity at
10 m, C , = 0.83 g l w , , C , = 0.83 g / 2 0 , , 9 is the angle between the wind
and the waves at the spectral peak, R, = U N u / ~ u , is the root mean
square surface deviation, and Y is the kinematic viscosity of water. Once
again, the parameters C,, C,, u, z, and z, can all be estimated from an
understanding of the wave climate (Melville, 1977; Donelan, 1980;
Donelan and Peirson, 1987) and depend on properties such as wind
duration, fetch length, and depth of water. Thus three major new
procedures for estimating neutral drag coefficients over a lake water
surface exist: from wave steepness, equations (3.41) and (3.42); from the
age of the wave field, equation (3.43); and from a knowledge of the wave
frequency spectrum, equations (3.44), (3.45), and (3.46).
The dependence on water surface state is extremely important for the
study of lakes as the fetch is usually limited, the winds are often diurnal
and the water depth is highly variable. However, there does not appear
to be available a study comparing the accuracy of any of the above
techniques and the choice remains mainly one of personal preference.
The scatter of data in measurements of CHNand CwNso far exceeds
the trends with changing wind speeds or sea state. Rayner (1981)
suggested using the constant value of C H N = CWN= 1.35 X based on
the works of Hicks (1972 and 1975) and Pond et al. (1974). Friehe and
Schmitt (1976) analyzed data from nine different experimental programs
for various stabilities. They found a value of 0.86 X for CH under
stable conditions and 0.97 X for unstable conditions. However,
Friehe and Schmitt (1976) also pointed out that if the large flux data from
Smith and Banke (1975) are added, the unstable coefficients become
1.46 x the value of CW was 1.32 X independent of stability.
By contrast, Large and Pond (1981) found good agreement with
similarity formulations for CDNand CwNbut found that C H N was equal to
1.13 x lop3 for unstable conditions and 0.66 for lop3 for stable layers.
The data analyzed by Smith (1980), when corrected for stability (see his
Figure 13), indicated a constant value of CHN= 1 X with perhaps a
slightly larger value for the unstable profiles. Anderson and Smith (1981)
found a mean value of CWNof 1.3 X but again with a large amount
of scatter. So far, the dependence on the wave climate parameters has
not been clearly demonstrated (Geernaert et al., 1987).
In closing, it is important to mention that McBean and Paterson
(1975), Bean et al. (1975), and Parker and Imberger (1986) have shown
334 Jorg Zmberger and John C. Patterson

that the wind field over a lake can be extremely variable: this has
enormous consequences for the behavior of the surface layer, as will be
seen in Section VI.

IV. The Surface Layer

The diurnal surface layer is the water counterpart to the atmospheric


internal boundary layer. However, it should be remembered from the
outset that the characteristic time scale h / u , with which the internal
boundary layer responds is of order 30seconds, whereas the same time
scale for the water surface layer is in the vicinity of 900seconds for a
layer thickness of 5m. The energy spectrum of the wind record used to
calculate the fluxes depicted in Figure 9 is shown in Figure 10; it is seen
that there is a thousand fold decrease in wind energy from 10-6s-' to

FREQUENCY (cyclesh)

Ro. 10. Power spectral density of the wind speed during the period shown in Figure 9.
Seventeen days of data were used to compute the spectrum.
Physical Limnology 335

10-3s-'. However, the energy at the higher frequencies is sufficient to


introduce considerable turbulent kinetic energy to the surface mixed
layer; since the period is comparable to the time scale of adjustment of
the water surface layer, it is unlikely that one would find, except perhaps
in the very near surface region, equilibrium profiles of the form given by
(3.14) to (3.16) (Dillon et al., 1981; Imberger, 1985; Muller and
Garwood, 1988). Even in the very surface region, it is unlikely that
equilibrium profiles exist because of the distortions introduced by wave
induced velocities. The only major exception would be in the case of a
strongly convecting surface layer, where the wind activity is small and the
time scale for the heat loss from the surface is set by the diurnal cycle or
by an even longer synoptic weather pattern. Such conditions have
recently been documented in lakes by Imberger (1985) and Brubaker
(1987) and in the ocean by Shay and Gregg (1984,1986).
Before proceeding with the discussion of the dynamics of the surface
layer and the associated vertical energy budget, let us consider some
actual examples of the evolution of the temperature field in a typical lake
situation. Recent microstructure measurements taken in the Wellington
Reservoir reveal the evolution of the diurnal surface layer. The data were
taken in February 1985 during an extremely hot period, when the
morning heating was strong and where, in the early morning, a southerly
wind with speeds up to 4 m sK1 prevailed which turned to a northwesterly
with a peak wind speed of 6 m s-l (at a height of 2 m) by 1300 hours. The
westerly remained active until 1900hours (see Monismith and Imberger
(1988) for a full description of meteorological conditions).
The initial profile from a region sheltered from the southerly wind is
shown in Figure lla-I. The surface (top 1 m ) had become strongly
stratified with a temperature differential of almost 1°C; the gradient
signal (Figure llb-I) shows little activity in the surface region except for
some isolated high gradient points (Imberger, 1988). The water below
this was turbulent, sustained by a weak gravitational underflow that had
been set up by the early morning differential cooling (Monismith and
Imberger, 1988). Ten minutes after the wind had moved north, causing a
gentle roughening of the surface, the temperature had changed to that
shown in the second profile. The temperature signal (Figure lla-11) and
the temperature gradient signal (Figure llb-11) show the formation of a
small mixed layer at the surface and also that the profile had become
active to a depth of 1.3m, even though the time since the surface had
become exposed to the wind was less than the time scale h l u , . The
336 Jorg Imberger and John C . Patterson

3
dT/dz (“C/m)

FIG. 11. (a) Examples of microstructure profiles collected at the entrance of Salmon
Brook, Wellington Reservoir on 26 February 1985. See Figure 26a for station locations.
Temperature:
I. 0928 SB20
11. 0939 SB20
111. 1002 c10
IV. 1021 c10
V. 1057 C10
VI. 1726 SBlO
(b) Microstructure temperature gradient signals corresponding to data in Figure 1la.
Physical Limnology 337

turbulence was established at small scales and was not the result of a
large overturn as has been normally associated with turbulence in a
stratified fluid (Gregg, 1987). Imberger (1988) attributed this rapid
spread of the energy to the leakage by internal wave action from the base
of the small surface mixed layer. Twenty minutes later, and at a
somewhat more exposed site further away from the shore, the surface
layer had deepened to about 0.3 m and the temperature distribution
within the lake was more uniform (Figure lla-111). There is clear
evidence of active turbulence extending below the base of the surface
layer to almost 2 m (Figure llb-111). This deepening process continued
and the profile taken another twenty minutes later reveals a “uniform”
surface layer extending to nearly 0.7m. However, as described by de
Szoeke and Rhines (1976), Imberger (1985), and Spigel ef al. (1986), the
interface at the base of the surface layer, initially sharp and sustained by
the surface-introduced turbulence, is now smeared by billow activity
energized by the momentum of the surface layer, which is clearly
illustrated in Figures lla-IV and llb-IV. The leakage through the base of
the mixed layer remained and the turbulent activity extended to nearly
2 m.
During the measuring period, the lake was exposed to a very strong net
heating flux of around 750 W m-*. At approximately 1030 hours, the
wind intensified, reaching a peak of nearly 6 m s-l at about 1100 hours. A
well-defined diurnal thermocline formed at 1.2 m, reflected both in the
temperature (Figure 1la-V) and the temperature gradient signal (Figure
llb-V), but two features are noticeable. First, the heat flux introduced
sufficient buoyancy at the surface to prevent a completely mixed surface
layer from forming, but the temperature gradient signal shows strong
turbulence in this surface layer. Second, the temperature gradient signal
continued to show turbulence below the base of the surface layer,
extending to about 2.0 m.
The wind strength remained at about 4-6 m s-l for the remainder of
the day and only began to decrease at 1700 hours. At the time the last
profile was collected, the wind had decreased to about 3ms-’. As seen
from Figure lla-VI, the surface layer was not well mixed, even though
the turbulence obviously extended down to 2.3m (see Figure llb-VI).
The gradient signal shows an intermittent, patchy turbulence within the
surface layer, although there was a well-defined wind-stirred near-surface
region extending down to 0.06m. The water column below 2.2m
appeared to have become quiescent.
338 Jorg Imberger and John C. Patterson
a b
I I I 1 I

I I I 1 I I
L2:.5 ’ 22.0 22.5 23.0 -50 0 50

TEMPERATURE (“C) dT/dz (“C/rn)

FIG. 12. Temperature microstructure data from the Wellington Reservoir (13 March
1982) at the central basin after a very strong sea breeze had deepened and tilted the surface
layer. (a) Temperature. (b) Temperature gradient. (After Imberger (1985).)

Now consider an example (Figures 12a and 12b) from the diurnal
sequence documented in Imberger (1985). The profile was taken at a
time when there was a net heat loss of about 250Wm-’, a strong
wind-induced surface layer velocity of about 0.12 m s-’, and an active
wind stress. As seen in Figure 12, the surface layer was extremely
well-defined (variation less than O.O3”C),the base of the mixed layer was
diffuse due to billow activity (Imberger, 1985), and the water below the
base of the surface layer was relatively quiescent.
The third example, shown in Figure 13, also comes from the diurnal
data set presented in Imberger (1985) but was collected when natural
convection completely dominated the surface layer dynamics. The profile
is characterized by a well-mixed surface layer with obvious thermals
Physical Limnology 339
a b

TEMPERATURE (“C) dT/dz (Qm)

FIG. 13. Temperature microstructure data from a penetrative convective period,


collected in the Wellington Reservoir on 14 March 1982. (a) Temperature. (b) Temperature
gradient. (After Imberger (1985).)

falling through the full extent of the layer and impinging on a sharp
interface.
In summary, these examples show that it is difficult to define the
surface layer in terms of the temperature profile. It is better to define the
layer as that depth of water directly energized by the surface fluxes and
the mean shear of the surface layer. The data also show that the
temperature, and thus the density, is rarely absolutely uniform in the
surface layer, especially when there is strong surface heating or when the
layer is retreating (wind stress reducing with time). Lastly, the surface
fluxes not only introduce turbulence into the surface layer but also, via
leakage through the base of the surface layer, into the deeper waters
making the above definition of the surface layer somewhat ambiguous.
Some of this surface-introduced energy energizes the water column as a
340 Jorg Irnberger and John C . Patterson

whole, but the exact fraction has not been documented in field
measurements.
The overall dynamics of the surface layer is determined by the
magnitude of the surface Wedderburn number (Imberger and Hamblin,
1982) given by (2.5). For W>>1, tilting of the isotherms due to the
applied wind stress will be small and horizontal variations are negligible.
This corresponds to strong stratification, light winds, and slow deepening
of the mixed layer dominated by surface-introduced turbulence reaching
the base of the mixed layer and eroding the interface. For W ((1,
deepening is dominated by internal shear production and occurs on a
time scale much shorter than horizontal convection in the surface layer.
This leads to a sharp interface downwind and a broad upwelling at the
upwind end of the lake (Monismith, 1986). In the initial classification (see
Spigel et al. (1986)), this regime was misinterpreted as having a
horizontal interface due to the envisaged rapid deepening. This question
is addressed in the next section. For intermediate values of the
-
Wedderburn number, W 1, upwelling and horizontal mixing become
important (Spigel and Imberger, 1980). This was confirmed by the model
of Imberger and Monismith (1986), the laboratory finding of Monismith
(1986) and by the field data of Imberger (1985) and Strub and Powell
(1986).
In the case where W > > l , the processes are essentially one-
dimensional, and as shown by Spigel (1980), the layer deepening can be
uncoupled from any internal seiching that may exist. There is a long
history of interest in such a one-dimensional surface layer. Munk and
Anderson (1948) discussed earlier observations and presented a theory
based on an eddy diffusion coefficient, dependent on stability, which
enabled the prediction of the surface layer behavior. The central task in
understanding the surface mixed layer is to quantify the rate at which the
kinetic energy of the turbulent velocity fluctuations is converted to
potential energy within and at the base of the surface layer as denser
water from beneath the layer is entrained and then mixed into the surface
layer. The turbulence available for this mixing may be generated at the
surface through pressure work, shear production, and wave breaking
(Kraus and Turner, 1967; Kraus, 1977; Cavaleri and Zecchetto, 1987;
Muller and Garwood, 1988) or in the surface layer and the base by shear
production (Pollard et al., 1973; Imberger, 1985). The transport and
redistribution of the turbulent kinetic energy within the surface layer is
dominated by large scale or secondary motions if present. The most
Physical Limnology 341

popular concept is of a coherent, well-ordered Langmuir circulation


(Langmuir, 1938; Scott et al., 1969; Leibovich and Radhakrishnan, 1977;
Pollard 1977; Leibovich, 1977a, 1977b, and 1980; Leibovich and Paolucci,
1980; Leibovich, 1983). From the examples of surface layer development
discussed above, the presence of such organized cellular motions are
apparently not ubiquitous (see also Muller and Garwood (1988)) and
only few direct observations exist (Thorpe and Hall, 1982; Weller et al.,
1985; Smith et al., 1987).
The turbulent kinetic energy budget in the surface layer has so far
remained elusive. Because direct observations of the vertical turbulent
fluxes of momentum, mass, or energy are still lacking, one must rely
completely on indirect evidence from measurement of dissipation of
turbulent kinetic energy. Such measurements can be used to validate
integral models (Imberger, 1985), scaling laws (Kitaigorodskii et al.,
1983; Dillon et al., 1981; Shay and Gregg, 1986; Brubaker, 1987; Gregg,
1987) and estimates of the Reynolds stress by equating dissipation to
production (Dillon et al., 1981). Detailed comparisons of this kind have
led to better agreement in the lake cases than those in the deep ocean,
suggesting that better estimation of surface fluxes, the smaller variance of
the turbulent fluxes (smaller waves), and the weaker influence of
horizontal advection all contribute to this better match. Further, it
suggests that the problem lies in capturing the full variability with the
measurements rather than with turbulent kinetic energy budget assump-
tions. An example of the mismatch encountered in the ocean is given in
Figure 14 (from Muller and Garwood (1988)), where the observed
dissipation rates are considerably larger than the estimated surface input
of energy.
A greater understanding of mixed layer energetics has come from
one-dimensional models that simulate vertical processes but neglect all
horizontal variations. These models are in two categories. First, a
number of attempts have been made to obtain solutions to the fun-
damental equations of motions using some type of closure assumption.
Munk and Anderson (1948) may be viewed as a zeroth order model
(Andre and Lecarrere, 1985), whereas the models by Mellor and Durbin
(1975), Mellor and Yamada (1982), Findikakis and Street (1982a, 1982b,
1983), Edinger and Buchak (1983), Franke et al. (1987), Shih et al.
(1987), and Murota et al. (1988) are higher order schemes. Excellent
summaries of the applicability of such closure schemes to mixed layer
problems is given in the review by Rodi (1987). The range of length
342 Jorg Imberger and John C. Patterson

10-7 10-6 10-5 10-4 10-3


o . o o ~ - *, , , , , , , ,
~ , , , , , ,,.[ , , , , ,,,, , , , , , ,,,, ,q
, , , ,,

0.10

0.20
n
a
5
0.30
BURST 19

0.40

0.10

0.20

0.30

0.40

0.50

me-’
FIG.14. Two examples of dissipation as a function of depth taken in the ocean and
compared with the similarity scaling for turbulence produced by both wind stress and by
convection. (From Muller and Garwood (1987); data originally from Gregg, unpublished.)

scales characterizing the surface layer dynamics and the unsteadiness of


the turbulence makes the application of closure schemes difficult to
justify. For instance, it is now well known that in a penetrative
convection surface layer, the heat flux in the lower half of the layer is
against the mean gradient, a process that requires the application of a
high order closure model.
Large eddy simulation schemes fall into this same class of models,
but are even more computer intensive. The method has been applied
successfully to a homogeneous wind-driven reservoir by Ivetic et af.
Physical Limnology 343

(1986), but once again, in the presence of stratification, the diversity of


scales and the non-stationarity of the turbulence has prevented the use of
the method.
Second, integral formulations pioneered by Niiler and Kraus (1977)
continue to be developed (Atkinson and Harleman, 1983; Spigel et al.,
1986; Heathershaw and Martin, 1987). As these authors point out, the
simplicity of physical insight afforded by the integral approach is
adequate justification for pursuing it.
Shear production of turbulence at the base of the surface layer was
recognized by Pollard et al. (1973) as an important source of energy for
the deepening process. Niiler (1975), Zeman and Tennekes (1977),
Sherman et al. (1978), and others have included shear production,
together with surface stirring as originally advocated by Kraus and
Turner (1967), in their mixed-layer models. Observations by Price et al.
(1978) verified the importance of shear production for oceanic mixed-
layer deepening, while the measurements of Thorpe (1978) in Loch Ness
provided evidence of billowing accompanying strong velocity shear at the
base of the mixed layer. Spigel and Imberger (1980) introduced a
classification for mixed-layer deepening that predicted the conditions
under which shear production would become important in small to
medium-sized lakes. Their mixed-layer-deepening algorithm was based
on the model of Sherman et al. (1978) but did include a separate
calculation to simulate billowing. However, Spigel and Imberger (1980)
did not explicitly account for the effects of billowing in the turbulent
energy budget, and their parametrization of the turbulent energy budget
assumed (in common with Sherman et al.) that mixed-layer turbulent
kinetic energy was a fixed proportion of external energy input by wind
and surface cooling. The assumption of fixed proportionality implies that
mixed-layer turbulence adjusts rapidly to changing external inputs as was
shown above; this may not be valid for diurnal simulations in which
meteorological forcing is varying rapidly. In addition, in deepening by
convective overturn, Denton and Wood’s (1981) experiments have shown
that the assumption of fixed proportionality is valid only under certain
boundary conditions.
The role of billowing is not so straightforward. Consider two parallel
streams separated by a sharp interface across which a velocity jump AU
and density jump A p (with the heavier layer on the bottom) occur.
Theoretical considerations (cf. Drazin and Reid (1981), Nishida and
Yoshida (1984), and Lawrence et al. (1987)) show that the interface is
344 Jorg Zmberger and John C . Patterson

unstable, and in practice, Kelvin-Helmholtz billows are always observed


along such an interface (Thorpe, 1969). The time for billows to form,
grow, and then collapse into small-scale turbulence is relatively short, of
order Tb = 20 A U / g f (Thorpe, 1973; Chu and Baddour, 1984), where
g’ = A p g / p , (po is a reference density). Values of T, in lakes are of the
order of minutes or less (Spigel, 1978). The ultimate mixing that occurs
after the breakdown of billows is of more relevance here than the details
of billow formation, and we use “billowing” broadly to include this final
mixing. The net result of billowing is that the sharp interface is replaced
by a shear layer of thickness 6 across which density and velocity vary
continuously. Simple energy arguments (Sherman et al., 1978) show that
6 is proportional to A U 2 / g f ;on the basis of numerical and laboratory
experiments, Sherman et al. suggest that

S = 0.3 AU2Jg’. (4.1)


Such a configuration is stable, and if nothing occurs to increase A U or
decrease the thickness 6, no further billowing will occur. If AU grows
(possibly due to an increasing wind stress) or 6 decreases (possibly due to
mixed-layer deepening-see Figure 1la-V) , then billowing will occur
until a new stable configuration is achieved.
There is thus a fundamental distinction between mixing caused by
billowing and mixing caused by thermocline erosion. Billowing broadens
an interface more or less symmetrically about the point of maximum
velocity gradient and thereby reduces the density gradient between
layers. Billowing by itself does not cause thermocline erosion or produce
any net mixed-layer deepening. Thermocline erosion is a one-way
process: an upper turbulent layer grows thicker at the expense of a lower
nonturbulent, or less turbulent, layer by entrainment of quiescent fluid
into the upper mixed layer. There is no tendency for density gradients to
be weakened during thermocline erosion and in some cases they may be
sharpened. Sharpening of density gradients is associated with convective
penetration or other stirring processes.
In lakes, billowing accompanies mixed-layer deepening (Thorpe, 1978;
Imberger, 1985) yet both processes drain kinetic energy from the mean
flow shear AU to produce turbulent kinetic energy. In this sense both
processes compete for a given supply of mean flow kinetic energy
proportional to $poAU2 per unit volume. The processes complement
each other in that a fraction of the turbulent kinetic energy produced in
both is used in working against gravity to increase the potential energy of
Physical Limnology 345

the water column; i.e., both processes cause mixing. Billowing utilizes
mean flow kinetic energy that would otherwise be available to produce
mixed-layer turbulence for entrainment, but at the same time weakens
the density gradient at the base of the mixed layer. Billowing reduces the
energy required for further mixed-layer deepening, since the energy
required to entrain heavier fluid is roughly proportional to the strength of
the density gradient across the base of the mixed layer. There thus arises
a rather complex and unsteady interaction whereby mixed-layer deepen-
ing sharpens a gradient, making it unstable to shear so that billowing
occurs. Billowing weakens the gradient, which is then more easily eroded
by further deepening, which leads to further billowing.
One of the principal goals of the model by Rayner (1981), and later
Spigel et al. (1986), was to demonstrate that the interaction between
billowing and deepening can be successfully modeled by accounting
simultaneously for the energetics of surface stirring, shear production,
billowing, and variability in the store of turbulent kinetic energy.
In these simple one-dimensional integral models, it is assumed that the
surface layer may be approximated in the integral sense by three distinct
zones (Niiler and Kraus, 1977; Spigel et al., 1986): a comparatively thin
constant stress surface layer of thickness y (see Figure 15) where
turbulent kinetic energy is produced and then exported to the fluid
below; a uniform central layer in which part of the surface energy is used
to mix the fluid and is lost to dissipation; and a thin front of thickness

ATMOSPHERK: TRANSFERS

p s r/”L p
3
/fWlHL

?
V
Z-H
-
F

rJ SURFACE
DRIFT
ZONE
WELL MMED LAYER
S, (FULLY TURBULENT)

HYPOLIMNION

Definition sketch of the surface layer model. (After Spigel et al. (1986))
346 Jorg Imberger and John C. Patterson

6 at the base of the surface layer. This transition marks the pronounced
density jump described above as the diurnal thermocline. Here, the
remainder of the turbulent kinetic energy generated at the surface plus
that generated internally by shear and less that which is locally dissipated
or radiated downwards by internal waves (Linden, 1975; E and Hopfin-
ger, 1986; Imberger, 1988) is used to entrain underlying fluid into the
central layer above. Using the symbols defined in Figure 15, the
governing differential equations (Niiler and Kraus, 1977; Spigel et al.,
1986) may be written as

dS
-=
a-
- -(s‘w’), (4.3)
at az
ap - - - ( ap ’ w-
_ ’)+--, a aQ
at dz cpaz (4.4)
au
-=
a-
- -(u’w’),
at 3.2 (4.5)

where E = u’* + v ’ +~w r 2 ,Q is the short-wave radiation, and the simplest


form of the momentum equation (4.5) is used. Niiler and Kraus (1977))
de Szoeke (1980), Ivey and Patterson (1984)) Janowitz (1986), and others
have included the influence of longitudinal pressure gradient and
Coriolos force. The rationale for using such a simple balance is discussed
in Spigel and Imberger (1980)) Kranenburg (1984)) Monismith (1985))
and Strub and Powell (1986). Monismith (1985) shows that (4.5) can be
used until the base of the mixed layer develops a baroclinic tilt due to the
second mode seiche; the time to the first deceleration is TJ4) where is
the period of the second mode seiche (the seiche associated with the
upper surface layer).
The boundary conditions at the free surface assume continuity of fluxes
across the interface:
-
Momentum: - -u’w ’(H) = u:, (4.7)
Heat: pOCpB’w’(H)
-
= QL+ HL+ H s , (4.8)
Salt: - = WS,,-
-s’w’(H) -
(4.9)
Mass: +
p ’ w ’ ( H )= p O [ - a e ’ w ’ ( H ) @ ’ w ’ ( H ) ] , (4.10)
Physical Limnology 347

where W is the water loss at the surface, QL is the net long-wave


radiation, HL is the latent heat loss, Hs is the sensible heat loss, (Y is the
thermal coefficient of expansion, and B is the salt coefficient of volume.
Spigel et al. (1986) assumed that the turbulent heat and momentum
fluxes at z = 5 - 612 (the very bottom of the base of the surface layer)
are zero, so that
- - - -
elw' = s ' w ' = p'w' = u ' w t = 0. (4.11)

The leakage or transfer of turbulent kinetic energy through the base may
be written

(4.12)

As seen from Figures 11, 12, and 13, this is usually nonzero. Imberger
(1988) postulated that the mechanism for turbulent energy transport was
by internal wave propagation, away from the base of the surface layer
along rays. Linden (1975) estimated from laboratory experiments that up
to 50% of the turbulent kinetic energy budget arriving at the base of the
mixed layer seeps through the base. By constrast, E and Hopfinger
(1986) suggested that while internal waves were generated, energy
radiation generally did not affect the entrainment rate. Scaling suggests
that the leakage A L be parameterized by

AL = CLA3N3 (4.13)
where CL is a coefficient, A is the amplitude of the internal waves
generated, N is the buoyancy frequency below the base, and h is the
depth of the mixed layer. At present, there are no conclusive data that
allows the determination of the correctness of (4.13) or the magnitude of
CL,and for this reason, Spigel et al. (1986) assumed CL= 0. It is worth
remarking that since their model gave very good comparisons with
observed field data, the leakage term must be proportional to the flux of
turbulent kinetic energy arriving at the interface, which is seen by noting
-
that A E:n/N.
Spigel et al. (1986) followed Niiler and Kraus (1977) and integrated
(4.2) to (4.6) from the bottom to the surface of the lake using the
boundary conditions given by (4.7) to (4.12). Further, closure was
achieved by introducing a set of efficiencies for each of the source and
sink terms. Shear production in the base layer minus the dissipation was
348 Jorg Zmberger and John C . Patterson

approximated by
Uzdh 1 d U2dh 1 d
--+--(U?h)- cbh = Cs 2-+--(Uzh)], (4.14)
2 dt 12dt 2 dt 12dt
where Eb is the average dissipation in the base layer and Cs is the
efficiency of production of turbulent kinetic energy. Experimental evi-
dence that allows the evaluation of Cs is sparse. Sherman et al. (1978)
reanalyzed the experiments of Kato and Phillips (1969) and Kantha et al.
(1977) to arrive at a value of between 0.2 and 0.5. Imberger (1985)
presented field data giving a value of 0.24 and Spigel et al. (1986)
obtained best simulation results with a value of 0.2.
The turbulence introduced at the surface constant stress layer may be
parameterized (Kraus and Turner, 1967; Niiler and Kraus, 1977) as

(4.15)

where c is the velocity scale in the surface drift layer (see Figure 15), CN
is the efficiency of energy production at the water surface, and u* is the
water shear velocity.
The source of buoyancy flux due to surface heating may be written
(Deardorff, 1970; Zeman and Tennekes, 1977; Rayner, 1981) in the
form

w: = g ( h + !){
2
[
Ly Q ( H ) - Q ( 5 -
POCP
): ] + ( y e " ( H ) }
+ pwss- -
2*g
POCP
1"
5-m
Qdz. (4.16)

The dissipation in the mixed layer may be written (Spigel et al., 1986)

E,h = $ESn, (4.17)

where E, is the average dissipation in the mixed layer. This form was
introduced by Mahrt and Lenschow (1976); similar forms were suggested
by Garwood (1977) and Zeman and Tennekes (1977), as well as by the
results of Willis and Deardorff (1974). The alternative parameterization
used by Niiler and Kraus (1977), Sherman et al. (1978), and others
consists of introducing a set of efficiencies to reduce each energy source
by a certain fraction; an approach that does not allow for the storage of
Physical Limnology 349

turbulent kinetic energy. Rayner (1981) proposed a necessary final


closure assumption which states that a fraction of the turbulent kinetic
energy iCFEF3 in the mixed layer together with the available shear
production (4.14) is utilized at the base to spin up the turbulence and to
energize the buoyancy flux (neglecting any leakage). This may be
expressed as
+> U : dh
- + - - (IUd: 6 ) ]
-E,
F'
2
312
2 [ dt 6dt
dh 6d I d
Esdh
2 dt 2po
+
= -- - Aph -- -- (Aph)
[ dt 2dt
+ --
12dt
(ApS')]. (4.18)

Given these assumptions and carrying out the integration yields (Spigel et
al., 1986) the necessary set of equations for the unknowns U s , h, 0, , S, ,
6, and E,:
dEs
h-=-(C,+
dt
CE)E,'"+ W: + C&:, (4.19)

(4.21)

(4.22)

and (4.1) and (4.18).


The evaluation (Spigel et al., 1986) of the remaining three coefficients
CN= 1.33, CF=0.25, and CE = 1.15 follows directly from a careful
analysis of all available data (Kato and Phillips, 1969; Tennekes and
Lumley, 1972; Wu, 1973; Deardorff, 1974; Kamail et al., 1976; Mahrt
and Lenschow, 1976; Stull, 1976; Willis and Deardorff, 1974).
Recently, there has been a renewed interest in the entrainment
problem described by (4.1) and (4.18)-(4.22), and there is greater
awareness that there are two distinct sources of energy for entrainment at
the base of the surface layer as expressed by (4.18),
dh I d
'[
fCFEfi2 and IC U : - + - - ( U Z S )
'
dt 6dt 1
Atkinson (1988) reviewed the literature for shear driven experiments
(see also Christodoulou (1986)) and concluded that the data support the
350 Jorg Zmberger and John C . Patterson

empirical entrainment law


1 dh
--= a - f3 Ri P;lR
(4.23)
U dt y + Ri
9

where a,f3, and y are constants, P, = U H / K , K is the coefficient of


species diffusion, and Ri = g'h/U2. Given that shear production domin-
ated the deepening process, (4.18) reduces to (assuming 6 to be small)

(4.24)

Comparison with (4.1) indicates that the depth of the mixed layer scales
the same as 6, which means that for a particular Us the depth is fixed and
cannot continue to deepen, contrary to (4.23). This was the basis of the
original Pollard et al. (1973) model, and may be explained by noting that
once a particular surface layer has a certain shear across its base, both h
and 6 will adjust until (4.1) and (4.24) are satisfied, after which time
there will be no further adjustment (Sherman et al., 1978; Lawrence et
al., 1987; Thorpe, 1987). The experiments of Narimousa et al. (1986) and
Narimousa and Fernando (1987) rely on a surface energy source u', that
will continually sharpen the interface and reduce 6, and the entrainment
law (4.23) therefore depends on the type of experiment and must be used
with caution.
This is best seen by substituting (4.22) into (4.24) and noting that g'h is
constant for no net heat flux, which then yields
1 dh Cy"
where Ri = g'hlu:. (4.25a)
u, dt -Ri'"'

Simple grid experiments contain no shear so that (4.18) reduces to the


relationship
1 dh - CF (4.25b)
E,ln dt 1 + g'hIE, '
Hannoun et al. (1988) have shown that the energy flux that arrives at the
interface is the same with and without the interface being present, and is
given by u3, where u is the root mean square velocity, which was also
observed to decay inversely proportional to the distance from the grid. In
grid experiments, we should therefore expect C F , the fraction of the
mean kinetic energy flux that arrives at the interface, to be a function of
both the depth of the surface layer and the turbulent intensities. Ignoring
Physical Limnology 351

the spin-up term in (4.25) and defining ii as the mean (over the depth of
the mixed layer) root mean square velocity allows (4.25) to be rewritten
ldh
--== CF
(4.26)
ii dt Ri’
where

(4.27)

Nokes (1988) has done a thorough review of data from all available
grid experiments and concluded that the rate of non-dimensional
deepening ranges in proportionality from Ri-’ to Ri-’.75. In this context,
the results from Kranenburg (1984) and Murota and Michioku (1986b)
give convincing evidence that the non-dimensional deepening is propor-
tional to Ri-l. In all recent work, investigators have defined the
Richardson number Ri=g’l/u*, where 1 is the integral scale of the
turbulence at the position of the interface (but in the absence of the
interface) and u is the root mean square velocity of the turbulence
arriving at the position of the interface. Unfortunately, there are no data
available on the relationship between the depth of the mixed layer and
the integral scale for surface layers so we cannot take this comparison any
further. It would be constructive to carry out an integral analysis of the
grid configuration to determine the relationship between ii and u, and to
establish if (4.26) would lead to the relationship suggested by Nokes
(1988):
l d h - 0.15
u dt Ri-’.’ * (4.28)

This is now possible as all the necessary spatial variations can be derived
from Hopfinger and Toly (1976), Hannoun and List (1988), and Hannoun
et al. (1988). Last, it is important to reconcile, as discussed by Nokes
(1988), the difference in the measurement of the entrainment rate dhldt:
indirectly from a measure of the change of buoyancy in the surface layer
and directly from the propagation rate of the interface.
Murota and Michioku (1986b) also carried out an experiment where
they used grid stirring together with a convective energy source and
successfully compared their results with the model put forward by
Sherman et al. (1978) and embodied in (4.19), although their value for
C , was 2.9, not 1.33.
b
0.0 w
VI
h,
1.0 -
2.0 -

7.0 1 ,/ /
21.50 21.75 22.00 22.25 22.50 22.75 23.00 23.25 23.50 23.75 24.00 21.50 21.75 22.00 22.25 22.50 22.75 23.00 23.25 23.50 23.75 24.00
TEMPERATURE ("C) TEMPERATURE ('C)

C d
0.0 0.0

1.o 1.o

2.0 2.0

x
E
3.0
-E 3.0

4.0 4.0
+ +
w
0 5.0 x 5.0

6.0 6.0

7.0 7.0

8.0
21.50 21.75 22.00 22.25 22.50 22.75 23.00 23.25 23.50 23.75
-
24.00
830
..
21.50 21.75 22.00 22.25 22.50 22.75 23.00 23.25 23.50 23.75 24.00
TEMPERATURE ('C) TEMPEAATURE (%)

FIG.16. Comparison of measured (solid line) and predicted (dashed line) temperature profiles. The profiles b, c, and
d have been corrected for effects of advection. (After Spigel el ul. (1986).)
Physical Limnology 353

The model described by (4.11) and (4.18)-(4.22) was used by Spigel et


al. (1986) to simulate the diurnal cycle documented in Imberger (1985)
for a surface layer in the Wellington Reservoir in March 1982. The
24-hour period included strong heating in the early morning, severe wind
in the afternoon when the Wedderburn number decreased to below 0.02,
and intense penetrative convection from midnight to early morning.
Comparison from these simulations is shown in Figure 16. Given that
none of the coefficients C s , CN, CF, or CE were adjusted for the data
and also given that a period of heating, stirring, shear production, and
penetrative convection all contributed to the surface layer deepening, the
comparisons are remarkably good.
Recently it was pointed out (Dickey and Simpson, 1983; Stefan et al.,
1983) that the exact distribution of solar radiation q ( z ) is most important
to the dynamical response of the surface layer. From (4.16), the
departure from the linear absorption curve determines the mechanical
buoyancy flux, and it is this departure that contributes to the direct
influence on the dynamical response.
In summary, we have shown that the identification of the energy
sources in the surface layer has allowed, over the last few years, the
construction of successful simulation algorithms. The success of these
models, however, depends on an accurate description of the surface
fluxes, a realistic model for the momentum in the surface layer, and the
removal of changes due to horizontal advection. Last, the model allows
intercomparisons with the full range of laboratory experiments such as
natural convection, shear-generated turbulence, and grid-generated tur-
bulence, but as yet this comparison does not seem to have been carried
out. The leakage of turbulent kinetic energy through the base of the
surface layer remains to be parameterized.

V. Upwelling

When the surface wind stress increases, the isopycnals surface at the
upwind end and deepen at the downwind end (Wedderburn, 1912;
Keulegan and Brame, 1960; Blanton, 1973; Stefan and Ford, 1975; Spigel
and Imberger, 1980; Monismith, 1986; and many others). This phenome-
non is called upwelling.
A demonstration of upwelling is given by the longitudinal transects
taken in the Wellington Reservoir (Figure 17a) in August 1988. The
354 Jorg Imberger and John C . Patterson

isopycnals prior to the start of a wind event are shown in Figure 17d; the
wind event started at about 1000hours and lasted for about 6hours
(Figure 17b), with a direction of 320" (Figure 17c), the direction of the
transect. The isopycnals after 4 hours are shown in Figure 17e, indicating
a high degree of disturbance. The hypolimnion was also disturbed but
retained a degree of horizontality. The upwind surface layer showed
strong upwelling with a well-defined horizontal gradient supported by
cold water brought up from the metalimnion.
The idealized density profile assumed by Heaps and Ramsbottom
(1966) and later used by Spigel and Imberger (1980) in their lake
classification scheme consisted of only two layers. This two-layer ap-
proximation was used by Thompson and Imberger (1980) and Imberger
and Hamblin (1982) to formulate the Wedderburn number (2.5) as the
nondimensional parameter determining the response of the surface layer
to an impulsive wind stress.
The initial value problem for the two-layer situation was solved by
Spigel(l980). For the case with rotation, the reader is referred to Csanady
(1982), Heaps (1984), Kielman and Simons (1984), and Horn et al.
(1986); the influence of rotation will not be discussed here. From Spigel
(1980), in the first stages of development of the surface layer following
the startup of the wind stress, the surface layer moves according to
(4.22). This surface layer velocity was shown by Imberger (1985) to be
applicable until the interface tilt reaches the center of the lake and
induces a baroclinic pressure gradient, which retards the motion. The
period of the surface layer seiche was given by ( h , < H)

T,= (5.1)

where L is the length of the lake, Ap is the density difference across the

FIG.17. Upwelling as recorded along the central channel of the Wellington Reservoir in
response to a strong northwesterly wind. Data collected on 24 August 1988. (a) Location
map showing plane of projection and measuring stations. (b) Wind speed as recorded in the
central valley of the Wellington at 1.4m above the water surface. (c) Wind direction
recorded as above. (d) Isopycnals at 0930, before wind had commenced. (e) Isopycnals at
1330 extending from the Gervase River to the main basin. (f) Example of a temperature
microstructure record, collected during the strong upwelling event shown in Figure 17e. (g)
Corresponding temperature gradient. (h) Corresponding dissipation as computed from the
Batchelor spectra fitting technique (Bar graph) and the Wigner-Ville technique (sticks).
23 000

22 000

21 000

20 000

-p 19000
GI
c
Y
LINE OF PROJECTDN
z l8OW AND WIND DIRECTION

:
8 17000
WELLINGTON RESERVOIR
2
s 16 000

15000

14 000

13 000

12 000
14 I 15000 16000 17000 18000 19000 20000 21000 22000 23000
LOCAL CO-ORDINATES Iml

-ur
-
E
0
6

L
fn
4
0
f 2
a
360

0
E
'f 270

:a:
k

0
g 180
5

90 u . .
I . ,. . . . .
2 '.O 237.5 238.0 238.5 239.0
DAY
356 Jorg Imberger and John C. Patterson

0.0
d DENSllY (kg/m3

2.0 --

4.0 --

6.0 --
E
E 8.0 --

12.0
’O’O 1
3000.0 4000.0 5000.0
PROJECTED DISTANCE (m)

0.0

2.0

4.0

--E
6.0

Eul 8.0
0

10.0
\
12.0 /*,*
‘+AO
o \ ,~
14.0
205
A
C60
A
C50
L
a2 /
A
’& A
2000 0 3000 0 4000 0 5000 0
PROJECTED DISTANCE (m)

FIG. 17. (conrd.)


Physical Lirnnology 357
2
358 Jorg Imberger and John C. Patterson

two-layer idealized density structure, h , is the depth of the surface layer,


h2 is the depth and p the density of the lower layer, and H is the depth of
the lake. Throughout it was assumed that the theory was applicable for
times up to T,/4 when the base of the surface layer begins to tilt. The
theory developed by Spigel and Imberger (1980), which provided the
background for the Wedderburn number concept, relied completely on
the two-layer model. However, it is apparent from the work of
Monismith (1986) that this model does not explain the upwelling that
already occurs at Wedderburn numbers considerably larger than one.
Monismith (1989, using the modal decomposition technique de-
veloped by Lighthill (1969) (see also Gill and Clarke (1974) and Csanady
(1972)), solved the n-layer case for wind stress varying arbitrarily in time.
He applied this general theory in three-layer form to the Wellington field
data reported by Imberger (1985). From the analysis, he showed that the
shear across the surface layer base AU,, was given by

which is the generalization of (4.22) to the case of an arbitrary number of


layers and where tl, is the displacement of the surface layer base
(between layer 1 and 2), A p 1 2is the density difference between layer 1
and 2, and h , is the thickness of layer 1 (the surface layer). The first term
on the right side of (5.2) is the pressure gradient induced by the slope of
the interface Cl2.
The importance of Monismith's (1985) analysis is that when hl and
A p l z are small (that is, W is small), mode 2 is excited, leading to a large
shear across the base of the surface layer with an associated strong tilt of
the interface f 1 2 ; the second interface 523 remains essentially horizontal
(Figure Ma). Conversely, where Ap,, is of the same order as Ap23 and
the top and bottom layers are of comparable size or larger than the
middle layer, the response is strongly influenced by the first mode and the
second mode with a comparable tilt in cl2 and f;23 and the introduction of
a large velocity in the bottom layer. By taking a range of surface layer
depths as measured in the Wellington field experiment, Monismith (1985)
obtained excellent comparisons between his theory and experimental
results for the thermocline tilt and induced shear; the three-layer
stratification example is a satisfactory model of the majority of lake
stratification patterns.
Physical Limnology 359

A-
b

FIG. 18. Schematic of the modal response for a three layer


. system.
. (a) Second mode
response: W small, L , large. (b) First mode response: W small, L, small.

The results from this model have two profound implications:


(a) A mode 2 response (Figure 18a) concentrates the shear, and thus
induced turbulence, in the upper surface layer. The strong tilt induced at
the base of the surface layer introduces active upwelling at even quite
moderate Wedderburn numbers (Monismith, 1986). Further, the lower
layer remains essentially stationary with c23
remaining horizontal.
(b) A mode 1 response (Figure 18b) introduces motion throughout the
lake, with tilting isopycnals in both the metalimnion and the hypolim-
nion. This motion leads to active turbulence throughout the water
column.
It is important to categorize a lake’s response as either mode 1 or 2. It
is clear from the theory underlying the Wedderburn number W that a
mode 2 response is induced when W becomes small (less than 10); the
360 Jorg Imberger and John C. Patterson

tilting of the base of the surface layer becomes stronger with decreasing
W (Monismith, 1986). The three-layer model has been used successfully
to explain why upwelling is induced for moderately small Wedderburn
numbers. With increasing wind stress relative to the buoyancy influence,
W decreases and the amplitude of the mode 1 response increases, causing
the metalimnion (layer 2) to tilt, which induces upwelling not only from
the metalimnion but also from the hypolimnion. However, it is not the
Wedderburn number that determines the mode 1 response, but the Lake
number LN (2.12); LN is the mode 1 counterpart to W in mode 2. This
may be illustrated by examining in more detail the experiments of
Monismith (1986). The stratification used in these experiments is well
approximated by a three-layer stratification with a constant density in
layers 1 and 3 and a linear transition between the two. For a rectangular
basin of length L, width B and total density difference between the top
and bottom layers of Ap, the Lake number takes the form

where

and h , , hz , and h3 have the same definition as above.


The results from Monismith’s (1986) Experiment 8 (Figure 19a-d)
show that a mode 2 response was induced with the base of the
metalimnion remaining essentially horizontal, but with the surface of the
metalimnion developing distinct upwelling very similar to that in Figure
17c for the Wellington Reservoir field data. The value of W was 3.9 and
that of LN was 6.4, indicating that the base of the surface layer was rela-
tively susceptible to the mode 2 response, but that the overall stratifica-
tion contained a large amount of potential energy. By contrast, the values
for Experiment 22 were W = 0.4 and LN = 0.15; the value of W had
decreased by a factor of 10 but LN had decreased by as much as a factor
of 50, indicating that in this experiment both the base of the surface layer
and stratification were relatively weak and could not resist the overturn-
ing moment supplied at the surface by the belt drive (Figure 19e-h).
These experiments must thus be reinterpreted as both W and LN were
alIowed to vary. The low Lake number value in Experiment 22 explains
why the boundaries of the metalimnion remained almost parallel at even
361

I
0.0-

O ' O P
5.0 -

-
-b

I
I 10.0-
n.
I-
w
n

15.0-
I , I I I ~o,ol-~---ll,oo~

20.0- I I I I I
50 100 150 200 250 300 50 100 150 200 250 300
DISTANCE (cm) DISTANCE (cm)

0.0 0.0

5.0- 5.0-

-E
x

--5
F
a
w
0
i
w
O

15.0- 15.0-

20.0 I I I I I T 20.0 I I I I I

FIG. 19. Contours of density field measured by Monismith (1986) in a laboratory tank.
The stress was introduced at the bottom with a belt moving from left to right.
(a) Experiment 8: W = 3.9, L , = 6.4, t = 0.32 T, .
(b) -
. , Experiment 8: W = 3.9, L , = 6.4, I = 0.64 T, .
(c) Experiment 8: W = 3.9, L , = 6.4, = 0.96T,.
(d) Experiment 8: W = 3.9, L , = 6.4, = 8.0 T , .
(e) Experiment 22: W = 0.4, L , = 0.15, = 0.35 T, .
( f ) Experiment 22: W = 0.4, L , = 0.15, = 0.53 T, .
(9) Experiment 22: W = 0.4, L , = 0.15, = 0.70 TI.
(h) Experiment 22: W = 0.4, L , = 0.15, = 1.05 TI.
362 Jorg Imberger and John C. Patterson
6 f
0.0
1

50 100 150 200 250 50 100 150 200 250


DISTANCE (crn) DISTANCE (cm)

50 100 150 200 250 50 roo I50 200 250


DISTANCE (cm) DISTANCE (cm)

FIG.19. (conid.)

extreme deflections (Figure 19f). Indeed, the whole metalimnion sur-


faced, leading to a direct mixing of water from the bottom and surface
layers. Experiments in which LN is kept large while W is reduced have
not been carried out, but field data from Imberger (1985), in the initial
phases of the wind build-up, had a Wedderburn number of 0.02 and a
Lake number of 12. The isotherms (isopycnals) below the base showed
essentially no tilt throughout the wind period. The Lake number
decreased to a value of 1.4 at the time of peak wind speed, and the data
indicated some tilting of the isotherms and some mixing below the
Physical Limnology 363

surface layer. By contrast, the Lake number for the data shown in Figure
17 was considerably smaller (0.24), explaining the greater mixing
observed at depth (see Section X ) .
W and LN are separate indicators, and for the case where W is small
but LN is large, only the surface layers respond to wind stress. Where W
and LN are small, the lake as a whole responds and we may expect that
vertical mixing will greatly increase throughout the lake.
The response in the case of continuous stratification (Monismith, 1986)
can now be interpreted in terms of these two nondimensional numbers.
For a linear stratification that reaches the surface, the mixed layer depth
is zero, as is the density jump across the base of the surface layer
(continuous stratification may be viewed as the limit where both
approach zero). This would mean that W is zero but LN is finite. The
expected response is a strong tilt of the surface isopycnals with active
upwelling, but at depth, the isopycnals remain horizontal. This case was
recently solved by Monismith (1987), using the same decomposition
technique mentioned above and assuming that the horizontal velocity
profile u(x, z, t) and the force field f ( x , z , t) could be decomposed as
follows:

n=l

where &(z) are solutions of the long wave form of the internal
eigenvalue problem (Gill, 1966) such that

and u,, fn, and 5, are the amplitude functions. Comparison with
experimental results from Monismith (1986) indicate that the fundamen-
tal mode was predicted well, but higher orders, which depend strongly on
the exact nature of the assumed stress distribution within the water
column were not as well predicted, although qualitatively, the agreement
was good.
So far the discussion has concentrated completely on the response of
times up to the quarter internal wave period for either mode 1 or 2.
Beyond this time, the internal displacement degenerates into seiches ( W
large, LN large) or damped motions (W and LN small). The damping of
the oscillations is caused by internal and boundary turbulence and we
364 Jorg Imberger and John C . Patterson

shall postpone the discussion of these processes to Section X, where


turbulence is reviewed.
For times longer than the damping time, it is possible to estimate the
deepening rate due to entrainment at the surface layer base for the case
where LN is large and W is small. Imberger and Monismith (1986) used
the results from Keulegan and Brame (1960), Kranenburg (1985), and
Monismith (1986) to formulate a model of a steady upwelling resulting
from a mode 2 response. Under such conditions, the experimental data
and the initial value problems discussed above lead to the following
conclusions:

(a) The interface tilts and opens at the upwind end over a period of
time approximately equal to the internal seiche period, (given
by (5.1)). The data suggest that the isopycnals are almost
horizontal at the bottom of the interface but slope considerably
upwards at the top of the interface, establishing a weak horizontal
density gradient in the mixed layer by a combination of upwelling
and horizontal variation in turbulent entrainment. The internal
circulation within the surface layer causes a general upwelling and
divergence of the flow at the upwind end, leading to a diffuse
interface. At the downwind end, there is downwelling, causing a
convergence at the interface, leading to an extremely sharp,
well-defined entraining interface.
(b) The velocity profile in the surface layer is such that the velocity at
the stress surface is approximately 20u, downwind (Kranenburg,
1985; Monismith, 1986) and about 224, downwind at the interface.
The flow at the interface was observed by Monismith (1986) to be
jet-like, concentrated a very short distance above the interface and
contained by a weak density gradient there. This differs from the
assumption made by Spigel and Imberger (1980) that the recircula-
tion flow in the mixed layer could be neglected for W > 1.
(c) The upwelling region is confined to the upwind end of the basin.
The density in the surface layer increases with time; for T > T,,
where T, is the characteristic longitudinal mixing time, the density
profile in the epilimnion varies linearly with distance along the
length of the surface layer.
(d) The net entrainment rule which best fits the experimental data
from Keulegan and Brame (1960), Kranenburg (1985), and Moni-
Physical Limnology 365

smith (1986) may be written

dh 1
-= C l u e Ri-', (5.7)
dt
where C1 is a constant between 0.07 (Kranenburg, 1985; Moni-
smith, 1986) and 0.23 (Wu, 1973).
These conclusions permit the construction of a simple model of
upwelling and mixed layer deepening for reservoirs where W 1 and -
L N>> 1 (Imberger and Monismith, 1986, Figure 20).
For this case, as the stress is applied to the free surface, the interface
tilt is set up by the mode 2 response so that the upper isopycnals take on
an angle given by the dynamic balance developed by Wu (1973), and the
boundary layer thickness is proportional to the displacement of the upper
isopycnals

where 6 here is shown in Figure 20, gh = ( p , - p o ) g / p o , p1 - po is the


initial value of the density difference at the base of the surface layer, hl is
the initial surface layer depth, and 5 is the mean interfacial deflection.

I
STREAMLINESOF UPWELLINGFLOW

u UPWELLING REGION -
SHEARSTRESS
I

I\ \ \...
.....I.\ .................. .).
............................
......\I ....
.............
.I... ..-. ............
A,... ........
4 4

I ISOPYCNALS

I PZ
I
I I
x-0 XIL

FIG.20. Definition sketch for model of mixed-layer deepening due to upwelling. (After
Imberger and Monismith (1986).)
366 Jorg Imberger and John C. Patterson

The model assumes that the return flow of 2u, drags interfacial fluid with
it from the downwind end to the upwelling region. It is assumed that this
fluid, which can be visualized as being planed off the interface, enters the
upwelling region and then is distributed longitudinally by shear flow
dispersion. A simple mass balance would indicate that

where pa is the average density in the surface layer, gi = g(pl - pa)/po,


and h is the mixed layer depth at any time. It follows that

(5.10)

If 6 satisfies (5.8) then (5.10) produces a result consistent with (5.7).


Thus, the Ri-’ entrainment law can be arrived at without invoking
interfacial shear entrainment or entrainment energized by turbulence
imported from the surface; it is instead the consequence of upwelling.
The buoyancy flux (given by (5.9)) introduced by upwelling into the
main part of the surface layer at the upwind end is mixed downwind by
shear flow dispersion. Thus, the perturbation buoyancy field in the
surface layer g&(x, t ) = [p(x, t ) - po]g/posatisfies the diffusion equation
(Fischer ef al., 1979)
--
ag:,
- Ex- a2g:, (5.11)
at ax2 ’
where
E~ = Czhu, (5.12)
is the effective diffusion coefficient. The coefficient C2 depends on the
shape of the velocity profile and on the distribution and magnitude of the
vertical diffusivity E,. For the log profile, characteristic of two-
dimensional open channel flow, Elder (1959) derived a value of 6,
However, C2 will remain unspecified for the moment. The width I of the
end region is approximately hl and can be neglected (Cormack et af.,
1974).
A simple quantitative model of how upwelling leads to mixed-layer
deepening can be formulated by assuming that the mixed layer is initially
homogeneous and by modeling the effect of upwelling with an impul-
sively started, and later constant, buoyancy flux of strength B at x = 0.
Thus, the temporal and spatial variations of g& can be found by solving
Physical Limnology 367
(5.11) subject to the following boundary and initial conditions.

ag:,
- B
(0, t ) = - H ( t ) , (5.13)
ax EX

ag:,
- (L, t) = 0, (5.14)
ax
g:,(x, 0) = 0. (5.15)
The required solution can be found in Carslaw and Jaeger (1978) as

2 [ierfc{(2n~+ L - ~ ) ( 2 ( ~ ~ t ) l ~ ) - ’ }
m

gk(x, r ) = 2/3(t/&,)ln
n=O

+ ierfc{(2nl+ x ) ( 2 ( ~ ~ t ) ’ ” ) - ’ } ] , (5.16)
where ierfc(x) = {exp(-x’)/dn} - x erfc(x) and erfc(x) is the com-
plementary error function. Equation (5.16) is plotted in Figure 21 in
terms of the quantity r = &(x, t ) ~ ~ ’ * / 3 - ’ t - ~for
’ ’ different values of

0
- ,
‘1L.10

L\\
0.2 a4 0.6 0.8 1

FIG.21. Curves of the solution (5.16) for I‘=g;(x, t ) ~ i ~ f i - ’ t - as


” ~functions of x at
different values of qL= L / ( E , ~ )(After
’ ~ . Imberger and Monismith (1986).)
368 Jorg Imberger and John C. Patterson

qL = L/(&,t)lR.When comparing experimental data with this solution, an


allowance must be made for the fact that horizontal density gradients are
created during the initial period of set-up by upwelling and by shear-
driven entrainment, i.e. gb(x, 0) =f(x). This can be done by adjusting
the time origin either through addition or subtraction of an offset.
Ideally, this offset should be much less than the characteristic timescale of
the dispersion process
L2
(5.17)
To==E,-
In addition, since /3 depends on h, it is not really constant throughout the
experiment. The solution will only be valid for times somewhat less than
hlRi
T,=-, (5.18)
ClU*

the time required for the mixed-layer depth to double.


From the solution plotted in Figure 21, it can be seen that the effect of
the upwelling flux is not felt at x = L until t -0.11 TO at which time
&(L) = O.O4gb(O). The total density difference in the mixed layer
reaches a maximum at t = 0.25 To, remains constant until t = TO and then
begins to drop as the mixed layer “heats up”. The solution for g,f,,(O, t)
can be shown to be
gb(0, t ) =2Bt
112 -112
Ex Jc
112
, (5.19)
for times less than 0.44 To. Finally, g,!,,(O, t ) will equal to 81, when

- 4Id
t = Tf - (“,W.($)
c: . (5.20)

If t > 0.44 TO, the factor appearing in (5.20) will not be n/4.
Figure 22 shows the data from three experiments reported in Keulegan
and Brame (1960), from two experiments reported in Kranenburg (1985),
and from Experiment 8 of Monismith (1986). In reducing these data,
Imberger and Monismith (1986) carried out the following operations:
(a) C,was first chosen to be 0.07 and, for all except Experiment 8, C2
was chosen to be 10.
(b) The resulting “raw” curves were plotted.
(c) So as to best fit the data to the theoretical curves, the dimension-
less time was offset by an amount A x and C1 was altered.
Physical Limnology 369

r
-
'1L
10-

10:
102 10-1 1 I

4%)
FIG.22. Plot of values of (I'/qL)calculated from densities measured with correction for
the initial conditions and adjusted for a best-fit of data at two values of 5. The solid lines
are the theoretical solutions while the symbols represent:
Monismith (1986), W= 3. 9, At'=0.060, C,=O.O7, C,=6;
A Keulegan and Brame (1960), W = 3.0, AT' = 0.015, C , -0.11, C , = 10;
0 Keulegan and Brame (1960), W = 2.4, AT'= 0.025, C , = 0.11, C , = 10;
0 Keulegan and Brame (1960), W = 1.5, AT' = 0.030, C, = 0.05, C , = 10;
+ Kranenburg (1985), W = 3.7, AT' = -0.015, C , = 0.11, C , = 10;
x Kranenburg (1985), W = 2.3, AT'= +0.010, C , = 0.11, C , = 10.
(After Imberger and Monismith (1986).)

The last step was interactive and performed first for the data from the c
position, say el, c
closest to = 0. Once a satisfactory fit was obtained for
el, the data at a second position, c2
> 0.5, was reduced using the final
values of C1 and AT. From trial and error, it became apparent that for
each experiment analyzed, a best set of values of AT, C1, and C2 existed
that appeared to minimize the differences between theory and measure-
370 Jorg Imberger and John C. Patterson

ments for both values of t;. In all cases, reasonable values of all three
parameters were obtained: the time offsets were always less than 0.04 To,
and the entrainment coefficient C, was between 0.05 and 0.133, well
within the range reported in the literature and cited above. The
difference between values of C2 chosen for the belt-driven flow and the
wind-driven flow may reflect differences in the structure of those flows,
especially in the way the shear stress and surface velocity vary with x .
Recently Imberger and Spigel (1987) have carried out measurements in
Lake Rotognaio in New Zealand during conditions of strong heating,
weak winds, and very turbid water (active algal bloom reduced the light
extinction depth to 0.5 m). The general upwelling circulation remained,
but a strong stratification persisted within the whole surface layer,
inhibiting turbulence and eliminating the entrainment surface. The
surface buoyancy flux outpaced the mixing caused by wind-induced
circulation. This type of upwelling is characterized by W being very small
but L N very large, and the circulation appeared as that described by
Monismith (1986) for the continuously stratified case. The major
difference observed was that with the buoyancy flux present a steady state
circulation appeared to form quite rapidly.
The other important variation involves the case of a variable wind
field. If the wind changes direction (Dickey and Simpson, 1983) or speed,
gravitational adjustments lead to large intrusions that may degenerate
into solitons (Kao et al., 1978; Maxworthy, 1980) or undular bores
(Thorpe, 1974; Farmer, 1978; Smyth and Holloway, 1988).
In summary, the upwelling response of the epilimnion (diurnal
thermocline) is determined by the magnitude of W, and that of the
metalimnion by the magnitude of L N . Where W is small, but LN large, it
is possible to model the entrainment process at steady state as simple
recirculation, where the entrained fluid is swept along the interface into
the upwind upwelling region, then is mixed in the surface layer downwind
by shear dispersion. If the stratification is complicated and admits
numerous eigenfunctions of equal importance, then an equivalent num-
ber of nondimensional numbers become important.

VI. Differential Deepening

The wind field over a lake surface is rarely uniform. Mesoscale wind
variability (McBean and Paterson, 1975; Bean et al., 1975) causes
Physical Limnology 371

variation over large lakes, and wind sheltering by the surrounding terrain
is the major factor affecting smaller lakes (Parker and Imberger, 1986).
Wind variability has two major effects. First, in areas with relatively high
wind speed, evaporation, and thus latent heat transfer, will be greater,
introducing a horizontal gradient in the surface heat flux. This phenome-
non, called differential cooling (Imberger, 1982), is discussed in Section
VII. Second, higher wind speeds on the exposed parts of the lake cause
both a more rapid local deepening of the surface layer and a greater
introduction of momentum. Both must ultimately be averaged across the
lake surface by gravitational adjustments of the isopycnals. The variabi-
lity in the surface layer deepening is called differential deepening
(Imberger and Parker, 1985).
Data from a field investigation near the headland of Salmon Brook in
the Wellington Reservoir illustrate these three phenomena extremely
well (Parker and Imberger, 1986). The data were collected along a
transect extending from a station in a sheltered embayment, station W4,
to the center of the basin exposed to the wind, Station C10 (Figure 23a).
The morning was warm and there was no wind. Shortly thereafter, a
southerly wind began to blow (wind at C10 was 2 m s-' and at W4 was
1m s-I); the surface layer was warm in the sheltered part of the transect
with some noticeable cooling out towards C10 (Figure 23b). This
difference was due to greater evaporative cooling in the open parts of the
lake. The wind intensified sharply at around 1540hours (rose to 4ms-'
at C10, but remained less than 0.6ms-' at W4) leading to noticeable
deepening at C10 (Figure 23c). The wind continued strongly until about
2000hours, when it decreased markedly to less than l m s - ' . The
horizontal density gradient associated with the isotherms is shown in
Figure 23d, implying a further deepening than that at 1805 hours. After
this time, the isotherms progressively relaxed (Figure 23e) until at
2300 hours (Figure 23f), the isotherms were once again virtually
horizontal.
Parker and Imberger (1986) investigated the evolution of the water
profiles by using the surface layer models described in the previous
section. They showed that the local deepening laws could be used to
describe the observations if local wind conditions were used and the
effects of horizontal advection were accounted for, although this ap-
peared to be a relatively minor correction. The results from Parker and
Imberger (1986) suggest that to obtain an average deepening law, careful
averaging must be done over the surface of the lake.
372 Jorg Imberger and John C. Patterson

15000
MAIN BASIN OF

1420

14kOO 15100 15600 16b00


X COORDINATE (m)

b: 1553 TEMPERATURE ("C)

N M K C10

1.0

FIG.23. An example of differential deepening recorded in the Wellington Reservoir in


March 1984. (a) Location map showing the isotherms. (b) Isotherms before the wind had
started at either the central or the sheltered station. (c) Data recorded at 1800 hours-the
wind speed in the central station had risen to 4 m s-1, but in the sheltered corner, the wind
was still less than 0.6111s-'. (d) Data recorded at 1948hourewind speed had dropped
everywhere to less than l.Oms-' immediately prior to taking this data set. (e) Data
recorded at 2137 hours-wind at both stations was less than 0.5 m s-', thermal structure was
relaxing and intruding out into the central part of the lake. (f) Data recorded at
2310 hours-the wind remained calm during the whole period, allowing the thermal
structure to adjust. The heat flux was beginning to go negative, causing slight cooling at the
surface. After Parker and Imberger (1986).
Physical Limnology 373

c: 1805 TEMPERATURE (“C)


00

1.0-
-,..- *-.- .*.-

--E
2.0-
h
X I

N M WClO K C10

0.0
PROJECTED DISTANCE (rn)

d: 1948 TEMPERATURE (“C)

PROJECTED DISTANCE (rn)

4.0
100.0 200.0 300.0 400.0 500.0
PROJECTED DISTANCE

FIG.23. (conrd.)
374 Jorg Imberger and John C. Patterson

f: 2310 TEMPERATURE ("C)

4.0
100 0 200.0 300.0 4000 500.0
PROJECTEDDISTANCE (rn)

FIG.23. (conrd.)

The case just discussed was for conditions of moderate W and large
LN. In Figure 24a, we show a much larger transect extending along the
length of the Wellington Reservoir approximately six hours after a
westerly wind had begun. The speed of the wind was 2ms-' at
0600 hours, then rose steadily and peaked at 6 m s-l at 1430hours. The
direction remained quite steady at due west (see Figures 17b and 17c).
The dramatic two-dimensionality of the structure is observed in the
isopycnals in Figure 24a with severe deepening between Stations C10 and
C45 and again between Stations C80 and C90; both sections were more
exposed to the strong westerly. The deepening of the isotherms between
C10 and C45 was due to the setdown associated with upwelling at C10,
whereas the slope of the isopycnals between C40 and C60 was due to the
more sheltered nature of this part of the valley. Confirmation of the
general tilt west-east due to upwelling is given in Figure 24b, which
shows data from a west-east transect at Station C40. The water in both
exposed areas was warmer than its surroundings, which is distinct from
what is observed in Figures 23b to 23f. This indicates that advection was
dominating the deepening process (Imberger, 1985). Further, the average
Wedderburn number for the central basin surface flow at 0927 hours was
0.25, and LN was approximately 1, implying a much less stable lake as a
whole, with mixing expected throughout; the shear associated with
upwelling (4.20) was sufficient even at Station C40, close to the end of
a DENSWY (kg / m3) b DENSITY (kg I m3)
0.0. 0.0.

5.0 5.0.

10.0- 10.0-

-
E E
--
15.0. 15.0.
E
I
I-
n
y1
n l !k
20.0- 20.0-

25.0- 25.0-

30.0.
:loo,

2000.0
CAW Iy
4000.0
CB", "I" I
6000.0 8000.0
Ac;\y"ycL

10000.0 12000.0
30.0-
0.0
A

200.0
A A

300.0
*
4dO.O
.I
500.0
A

6bO.O
A

7[
EAST WE
CUMULATIVE DISTANCE (m)

FIG.24. (a) Transect along the Wellington Valley (see Figure 17a). The wind was coming from the west (as shown in Figures 17b and
17c) and blowing across the lake between the 2000 and 5OOO m mark as well as between the 12000-16000m mark. The isopycnals show
extreme differential deepening, the wind strength at the time was 6ms-'. (b) Cross section at Station C40, from east to west, the
differential deepening, shown in Figure 24a was accompanied by strong upwelling. (c) Temperature. (d) Gradient temperature. (e)
Dissipation as recorded with microstructure instrument at Station C40 at 0947 hours. Dissipation as measured by both the Wigner-Vie
technique (sticks) and the Batchelor spectrum fitting technique (bar chart).
376 Jorg Imberger and John C. Patterson
F
0 1
OL
1 I I I I
0 (D N
l- z P
N
(u)Hld30
Physical Limnology 377

the basin, to induce very active mixing throughout the water column (see
Figures 24c, 24d, and 24e).
De Szoeke (1980) addressed the problem of a variable wind stress over
the ocean. However, he aimed to determine the variability induced by
the Ekman pumping driven by the curl of the surface wind stress and not
by the differential rates of vertical entrainment. Maxworthy and Moni-
smith (1988) were motivated, in part, by the observations in the
Wellington (Parker and Imberger, 1986) and conducted a grid stirring
experiment where the grid extended over only part of a long, stratified
tank. Their results illustrated the sequence of events immediately after
the commencement of the grid oscillations. First, the turbulence intro-
duced by the grid mixed the underlying fluid and set up a downward
propagating entraining front immediately below the grid. Second, this
front continued to propagate downwards until either the turbulence
became weaker due to the greater distance from the grid (Hopfinger and
Toly, 1976), or the mixed fluid collapsed and intruded horizontally into
the neighboring quiescent ambient fluid. Third, as the intrusion propag-
ated out into the main tank, it reached a stage where the end of the tank
began to slow its progress, once again influencing the rate of descent of
the entrainment front immediately below the grid.
Using the various relationships for intrusion speed and entrainment
rates, Maxworthy and Monismith (1988) derived a relationship for the
speed of propagation of the entrainment front under the grid and
satisfactorily verified these relationships with the experiments. The grid
mechanism, while convenient, is a poor analogue for wind-induced
turbulence. Thus, even though the flow scenario will be the same, it is
more useful, as Maxworthy and Monismith (1988) did, to write down the
relationships for the case where the mixing is carried out by a wind stress.
Consider a strongly stratified surface layer over part of which a strong
wind stress is suddenly initiated with a shear velocity of u , . Under the
wind stress, the initial deepening will be given by (4.25b)

1 dh - C,
u , dt Ri* ’

which is the relationship relevant for a linearly stratified fluid and where
C , = 0.1 (Kranenburg, 1985). As shown by de Szoeke and Rhines (1976),
surface-induced turbulence and deepening dominates initially, until the
shear at the base of the mixed layer (5.2) builds up. For longer times,
378 Jorg Zmberger and John C. Patterson

shear production (4.25a) dominates and the entrainment rate is described


by:
1 dh U
-- = (2Ca)'n(-?)
u* dt Nh '
where C,= 0.2 to 0.3 (Spigel er al., 1986).
These deepening processes will continue (Maxworthy and Monismith,
1988) until the outflows/inflows become appreciable (see Figure 25) and
dilute the effort expended by the vertical mixing processes.
It is well established (Maxworthy, 1972; Imberger et al., 1976; Manins,
1976) that the intrusion can either be governed by an inertial buoyancy
balance or a viscous inertial buoyancy balance; in the reservoir, however,
the scales are such that invariably the inertial buoyancy force balance
prevails. This means that the intrusion thickness is given by (for a
two-dimensional intrusion)

and the speed by


c = 0.44(Nq)1'2,

where q is the two-dimensional volume flux and N the buoyancy


frequency. However, as observed by Maxworthy and Monismith (1988),
6 is not necessarily equal to the depth H of the entraining front under the

EXPOSED

- SHELTERED

~~

ENTRAINMENT
FRONT

STRATIFIED 4-

FIG. 25. Schematic of differential deepening.


Physical Limnology 379

grid. The flow out of the turbulent area is governed by a gravitationally


driven shear flow, which is retarded by the vertical turbulent transport.
Such flows have received extensive treatment in estuarine situations;
detailed solutions are described in Scott and Imberger (1988). The flow
itself was recently modeled in the laboratory by Linden and Simpson
(1986), who, however, do not appear to have connected the dynamics of
their flow with the pioneering work of Taylor (1954) on shear flow
dispersion (see Fischer et al. (1979) for a general reference). It is
therefore better to re-analyze the Linden and Simpson (1986) experi-
ments in light of the estuarine work and the analysis presented by
Maxworthy and Monismith (1988).
Suppose the wind introduces a vertical mixing coefficient given by
(Fischer et al., 1979)
E, = 0.06 u,h. (6.5)
The flow under the grid then may be derived, if it is assumed that the grid
is much larger than the mixed region is deep ( h / L< I), from the long
box convection problem, where L is the length of the box.
A simple balance between baroclinic vorticity generation and vertical
diffusion leads to the scale of the velocity under the grid (Imberger, 1974)

where g’ is the reduced gravity across the turbulent front and Lf is the
length of the density variation at the outflow end; this length increases
with time as the intrusion increases to bring outer fluid into the turbulent
zone. This balance may be expected to be valid provided Lf > h2ule,
(Fischer et al., 1979).
The effective longitudinal dispersion coefficient is given by Fischer et
al. (1979) for such a velocity profile:
h2u2
K, = 5.3 x
€2

and a good approximation to Lfis given by

L‘ = (Kxt)1’2 (6.8)
for times which are short enough so that L > Lf.Combining (6.5) and
(6.8) yields the relationships for both the dispersion coefficient and the
380 Jorg Zmberger and John C . Patterson

peak velocity:
(6.9)
12 3 114
u=1.2(g@)
u*t , (6.10)

where it was assumed that the depth h adjusts slowly compared to the
time it takes for the flow to adjust to the changing Lf.
Unfortunately, the data presented in Linden and Simpson (1986) are
not sufficiently detailed to allow comparison. Maxworthy and Monismith
(1988) balanced the deepening given by (4.23) with the vertical advection
induced by the outflowing intrusional such that

(6.11)

However, if we consider the example shown in Fig. 23c, then K x =


4x m s-l, so that in the time deepening has taken place, Lfhas only
grown to 6 m , far too small to influence the process; it is therefore not
surprising that Parker and Imberger (1986) obtained good comparisons
between the one-dimensional model and the field data once they used
local wind speeds. The oversight in the paper by Maxworthy and
Monismith (1988) was that no distinction was made between the length
scale Lf and L.
In summary, the concept of differential deepening appears to be well
established and documented in the field. The laboratory experiments by
Maxworthy and Monismith (1988) have given an explanation for the
connection between the deepening under the exposed areas and the
intrusional flows set up by the buoyancy imbalance introduced by the
wind. However, it appears that for diurnal cycles, intrusional flow is of
secondary importance to the deepening. However, the intrusional flows,
which can reach velocities of up to 5cm per second (Parker and
Imberger, 1986), are of primary importance in redistributing water
horizontally once the wind event is over.

Vn. DEerential Heating and Cooling


Field data collected in Salmon Brook, a side arm of the Wellington
Reservoir (Monismith and Imberger, 1988), provides a good illustration
of differential heating and cooling. Salmon Brook (Figure 26a) runs
Physical Limnology 381

north-south with a central uniform valley. The surrounding terrain is


hilly and the upper reaches of the side arm are sheltered from the wind.
Data were collected in late February 1985 during a hot (39"C), clear
day, where the peak net heat flux at the surface reached 960 W m-*. The
wind was less than 2 m s - ' all day except briefly between 1500 and
1830 hours, when a westerly wind with a speed of 5.5 m s-' sprang up.
The night had been relatively cool and still, with only a small northerly
breeze around midnight which reached 4 m s-'. This stopped completely
by 0700 hours.
An early morning transect along the central valley of Salmon Brook
from the upstream extreme to Station C10 (see Figure 26a) is shown in
Figure 26b. Cold water, which had formed around the perimeter of the
side arm during the night, fell under gravity from the surface to about
7 m . A transverse transect taken across SB20 (Figure 26c) shows the
symmetric plunging of the boundary water. Drogue tracking indicated a
drawing of central water into the side arm along the surface consistent
with the temperature observations that the night cooling had formed cold
boundary water, which downwelled around the perimeter of the side
arm. The inflow occurred over 50-70% of the depth of the underflow
(7 m), comparable to the laboratory results of Harashima and Watanabe
(1986).
In the early afternoon, after strong heating all day, the stratification
changed to that shown in Figures 26d and 26e, indicating strong boundary
heating around the perimeter with baroclinic pressure gradients of the
temperature field forcing the water away from the perimeter into the
main basin. However, Monismith and Imberger (1988) observed that
during this time the surface water was still slowly moving into the side
arm, holding the warm water near the perimeter.
By late afternoon, the surface water motion had reversed and a strong,
buoyant surface overflow (velocity approximately 0.10 m s-l) confined to
the top 1m was observed to be moving the water out into the main basin.
By 2200hours, the whole of the surface water in the top 2 m had been
replaced by this mechanism.
The westerly wind that sprang up at 1520 hours (see Figures 26f, 26g,
and 26h) immediately moved the thin surface buoyant plume onto the
east bank; the influence of the wind was progressively stronger towards
the mouth of the brook. It is interesting, in passing, to observe the
general shape of the upwelling pattern in Salmon Brook as one moves
382 Jorg Zmberger and John C . Patterson
a

1moo-

--
E
5
I -
14600-

> SALMON
14200-

N
14b00 15dOO 1 seoo 16hOO
X COORDINATE (m)

TEMPERATURE (“C) TEMPERATURE (“C)


C
0.0-

?‘
2.0-

,
4.0-

-E-E 6.0-
0

8.0-

10.0- 10.0-

12.0- - 0 . 21
I 200.0 4000 600.0 800.0 1000.0 12 0 0.0 100.0 200.0 300.0 400.0 51 0

PROJECTED DISTANCE (m) PROJECTED DISTANCE (m)

FIG.26. An example of differential heating and cooling, illustrated with data collected
in Salmon Brook on 25 February 1985. (a) Bottom bathymetry of the site showing site and
transect positions. (b) Longitudinal transect taken early morning, starting at 0849 hours. (c)
Transect across Salmon Brook at Station SB20, early morning, commencing at 0909 hours.
(d) Longitudinal transect illustrating midday heating, taken at 1333hours. (e) Cross-
sectional transect at Station SB20, showing thermal heating near boundaries, data taken at
1340 hours. (f) Transect taken at SB25 at 1639 hours, illustrating the influence of a very
weak westerly wind. (g) Transect taken at SB20 at 1536 hours, the wind at this station was
somewhat stronger. (h) Transect taken at SBOO, commencing at 1639 hours. This section
was almost completely exposed to the wind, which now had reached a value of 4 m s-’. (i)
The net heat flux variation over the five days of the experiment as computed by the rate of
change of heat of the total water column at Station SB30.
Physical Limnology 383
d TEMPERATURECC)

“-I 10.0

12.0
5830

2dO.O
--
4ob.O Sdo.0
A
----..._
- .-
8dO.O
L

PRQlECTED DISTANCE (m)


A
c10
;
lObO.0 12w.0

a TEMPERANRE (‘C) f TEMPERANRE (“C)

4.01 I 2.01
-..._ ,.-*.----.

10.0

s020w S020E S025W S025E


12.0
0.0 100.0 200.0 3W.O 400.0 500.0 0.0 20.0 40.0 60.0 80.0 100.0 120.0 140.0 1 1.0
PROJECTED DISTANCE (m) PROJECTEDDISTANCE ( m )

FIG.26. (cod.)

from the sheltered SB25 station to the more exposed SBOO (Figure 26h)
station, and to compare this with the discussion in Section VI. At SB25,
the Wedderburn number was around 5-10, as indicated by the very weak
surface upwelling and the almost horizontal isotherms immediately
below. At SB20, the whole diurnal thermocline was tilted, surfacing at
the west end of the side arm, indicating that the Wedderburn number was
around 1. Lastly, at SBOO, the Wedderburn number was calculated as 0.2
and the isotherms indicate a very similar pattern to that shown in Figure
19h. where the Wedderburn number was 0.4.
384 Jorg Imberger and John C. Patterson
TEMPERATURE (‘C)
g

SB2OW SBBOE
6.0
200.0 3W.O 400.0 500.0 600.0
PROJECTED DISTANCE (m)

TEMPERATURE(‘C)
h TEMPERATURE (‘C) i
0 0-
I

I
c

I
I

-
TOTALNETHEATFLUXAT

iI
10-

2 0-
-1000 MUOBSERVED
HEAT INPUT AT WSi
-
, .
m DAY 65
I

,
7
DAY 68
DAY 67
DAVU
O A V ~

-E v-
cc
’=5.
z30-
a -
- - - - - M- - -
U OQSERMD
Lu * 8URFACECOOLING
4 0-

+I00
5 0- 0 0600 1200 I800 2400 0600

SBlOE
A
l.,/= 0- *O-

./--_*..-,..
,,
.-.o-

A
SBlOW
A
TIME (H)

60
I I I

FIG.26. (conrd.)

The consequences for the heat budget of this cyclic flow pattern are
illustrated in Figure 26i. The data show that the water column at Station
SB30 (the head of the side arm) responded as if twice the surface heat
load had been applied in both the heating and cooling phases. This is
further strong evidence that the inertia of the water leads to a large phase
lag between the thermal forcing and the flow response.
Flow patterns associated with the cooling phase are driven by the
greater temperature drop around the boundary; it is yet to be established
whether a variable heat flux also contributed to the excess temperature.
Physical Limnology 385

On the heating phase, both the reduced depth and the reduced latent
heat loss (lower wind speed in the sheltered areas) around the perimeter
lead to the greater temperatures around the boundaries. Steady convec-
tion in a long, shallow cavity, an approximation of the side arm, has
received a great deal of attention for two-dimensional configurations
(Hart, 1972; Cormack et al., 1974; Imberger, 1974; Bejan and Tien,
1978; Bejan and Rossie, 1981; Bejan et al., 1981; Inaba et al., 1981;
Simpkins and Dudderar, 1981; Hart, 1983a, 1983b; Drummond and
Korpela, 1987) and in three-dimensions by Imberger (1976) and Scott
and Imberger (1988). These solutions rely on a viscous buoyancy balance
at first order and all have stable stratification that changes from one with
vertical isotherms at low Rayleigh numbers to one with horizontal
isotherms and boundary jets along the top and bottom boundaries at very
large Rayleigh numbers, such as those that operate in a side arm of a
reservoir. These solutions are of limited value, however, as they require
an imposed steady horizontal density gradient, with zero heat flux at the
upper and lower boundaries. Surface heat fluxes can be incorporated into
the solution technique (Cormack et al., 1975; Scott and Imberger, 1988)
but unless vertical velocities are allowed at first order, these heat fluxes
can only be incorporated at second order, making the solution techniques
less applicable.
Most importantly, in these diurnally cycled flows the inertia of the
water leads to a velocity field almost 90" out of phase with the thermal
forcing. In the data shown in Figure 26, the velocities did not reverse
until about 1500 hours, seven hours after the net heat flux had changed
from a loss to a gain (Monismith and Imberger, 1988). The explanation
of this observation lies in the following scaling arguments. A thermal
gradient is instantaneously imposed on a long cavity of fluid. This
gradient would induce a flow which would be governed by the balance

_
dP -
dz
- -gP,

leading to a time scale for spin-up;

T=-U (7.3)
Kgh'
386 Jorg Imberger and John C. Patterson

where u is a typical velocity of the final flow, K is the density gradient

and h is the depth of the flow. Inserting typical values derived from
Figure 26d leads to a time scale of two hours, which is reasonable given
the gradual build-up of the gradient.
The unsteady cavity problem has also been investigated by Patterson
and Imberger (1980) and Patterson (1984). The latter paper used internal
radiative heating to drive the flow rather than heating and cooling at the
vertical end boundaries. The critical parameter to emerge from the long
box problem is Gr A4, where

gcu ABh3
Gr = (7.4)
Y2 ’

A = -h (7.5)
L’

cu is the coefficient of thermal expansion, and At) is the temperature


differential. Once again, the applicability of these theories is marginal
due to the cyclic nature of the heat sources and the very different
boundary conditions. However, evaluation of the parameter P = Gr A4
leads to a value well in excess of 1 for a typical side arm, indicating that
the flow in side arms must be dominated strongly by inertia even if an
effective diffusivity Y of m2s-l is used, as would be indicated by the
turbulence measurements in these flows (Imberger, 1988).
The integral solutions advanced by Sturm (1981) and Jain (1982) seek
to explain the laboratory findings of Brocard and Harleman (1980). All
three deal with side arms with a strong surface heat loss, sustaining an
unstable upper water column and a longitudinal circulation driven by a
longitudinal temperature gradient. These investigations were motivated
by cooling pond configurations where there is a net surface heat loss at all
times and a steady state solution can be contemplated. Harashima and
Watanabe (1986) presented further laboratory verification that the
solutions found by Sturm (1981) and Jain (1982) meaningfully describe
these flows. However, since the surface temperature in a cooling pond is
generally above the normal equilibrium temperature for the prevailing
Physical Limnology 387

meteorological fluxes, a longitudinal surface temperature gradient always


develops. By comparison (Figure 26b) no such gradient develops during
the cooling phase, which is better described by a source of negative
buoyancy distributed around the perimeter of the lake. By contrast, the
heating cycle may be modeled by a line source of buoyancy in
combination with a variable heat flux along the length of the side arm.
Imberger (1985) pointed to a potentially important convection problem
in lakes: differential absorption. Light and short wave radiation penetrate
the water column to considerable depth; the exact depth is dependent on
the concentration of particulate matter and the color of the water (Kirk,
1983). Where the water is transparent, short wave energy is distributed to
considerable depth; conversely, in areas of high turbidity, radiation will
penetrate to a much lesser depth and more heat will be absorbed near the
surface of the water column. This causes the surface temperature in the
turbid area to rise, and the water immediately beneath, shaded by the
turbidity, will become colder than in neighboring areas. Large tempera-
ture changes may therefore be expected to build up over successive days
of heating where the turbidity distribution is uneven.
Data collected recently in the upper reaches of the Canning Reservoir
are shown in Figures 27a and 27b, for the transect from C45 to C240 (see
Figure 27c). The isotherms show that a strong surface warm water lens
had formed where the water was turbid near Station CA100. Underneath
this lens, the water was considerably colder than the water at the same
depth at Station CA45. At the time of these measurements, a weak
variation in salinity existed at the upstream end of the reservoir, enough
to be used as a tracer, but at a low enough concentration not to
significantly affect the density field. The corresponding salinity structures
are shown in Figure 27b. An intrusion had formed at a depth of 3 m and
was propagating downstream and upstream. At the surface, the hot water
lens was spreading out, as evidenced by the spread of the isohalines,
although the speed was considerably slower than in the subsurface
intrusion. These flows are driven by the weak density gradients induced
by differential absorption. The flow may be termed a thermal siphon, and
will in many cases be a major contributor to a horizontal redistribution of
water in areas of patchy turbidity, an important factor in the kinetics of
phytoplankton communities.
Two theoretical models have been developed from the results shown in
Figures 27a and 27b. Patterson (1984) used the long cavity with a
longitudinally variable absorption as a model. He assumed a longitudinal
388 Jorg Imberger and John C. Patterson

dependence of the radiation flux

Q(x, 2 , t ) = 2Qoh(x - $), (7.6)


where h is the depth of the cavity, L is the length of the cavity, and Qo is
a constant, the value of which was used to fix the intensity of the
radiative heating.
The flow was assumed to be laminar so that the thermal energy

b SALINITY (psfi)

DISTANCE (m)

FIG.27. (a) Isotherms (“C) along a transect from Station CA45 (distance = 4377 m) in
the main Canning River tributary to Station CA24D (distance=9147m). Data was
collected between 1704 hours and 1813 hours, 15 December 1983. For station locations, see
Figure 7b. (From Imberger (1985).) (b) Isohalines (pss) for the same data set as in (a).
(From Imberger (1989.) (c) Map of Canning Reservoir showing measurement stations.
Physical Limnology 389

POISON GULLY

CANNING RESERVOIR

CA180

CA230

oooo 1000 2000 3000 4Ooo 6000


LOCAL CO-OROINATES (m)
FIG.27. (contd.)

equation became
ae de ae
-+ u -+ W - = K V20 + Q(x, Z, t ) . (7.7)
at ax az
A scaling analysis revealed that the flow development, for the initial
value problem, depended critically on the magnitude of the parameter
P = Gr A4, (7.8)
390 Jorg Zmberger and John C . Patterson
where
gaQoh4L3
Gr = (7.9)
Y3

is the Grashof number for the problem and A is defined by (7.5). Eight
regimes of flow were defined. There are two extremes: P > Pr4 (where
Pr = Y / K , the Prandtl number) is the regime where inertia and convec-
tion dominate the final stages of flow. Second, P<Pr-'A4 gives a
thermal field described by a simple balance between conduction and
radiative heating.
Introducing typical values for the problem illustrated in Figures 27a
-
and 27b leads to the first regime (assuming Pr 1) as P is much greater
than one. For this regime Patterson (1984) observed a single gyre
centered at the transition between heating and cooling. The eddy was flat
and supported horizontal isotherms.
Trevisan and Bejan (1986) modeled the turbidity cloud as the perfect
reflector. This is the extreme in differential absorption and, once again, a
simple gyre was generated at the boundary between the absorbing and
reflecting areas although the gyre computed by these authors was not
dissimilar to that found by Patterson (1984). Neither model describes the
real situation; the absorption is differential in horizontal and vertical
directions and the patches are usually three-dimensional.
The original problem that motivated research into convective motions
was in cold lakes and concentrated on the thermal bar problems
(Rodgers, 1971; Spain et al., 1976; Mortimer, 1987). This phenomenon
occurs when a lake, initially at a temperature less than 4"C, begins to
warm. Similar to the observations described above, this warming occurs
more rapidly around the perimeter due to the shallow depth and the
warmer river inflows. In this sense, it is identical to what is believed to be
a good model for the data shown in Figures 26d and 26e. The extra
complication is that the equation of state is nonlinear at these low
temperatures, with a maximum density at 4°C.
Once the boundary fluid has a temperature greater than 4"C, it is
lighter and propagates along the surface. As it meets the colder (<4"C)
water offshore, mixing produces some 4°C water that sinks (Marmoush et
al. , 1984) and produces a complicated frontal mixing 'pattern, distinct
from the normal buoyant surface overflow which has a well-defined
frontal roller (Luketina and Imberger, 1987). Elliott and Elliott (1970)
and Ivey and Hamblin (1988) have carried out laboratory experiments in
Physical Limnology 391

long cavities and Ivey and Hamblin (1988) in particular present graphic
data for the formation of a strong region of convergence, where the 4°C
water plunged in a plumelike sheet to the bottom of the tank. By analogy
with the long box problem, Ivey and Hamblin (1988) suggest the thermal
bar (the region of strong convergence) has a width of O(h Ra-"4), where
gtu( A q 2 h 3
Ra = 9 (7.10)
YK

and
p = po[ 1 - a(8 - 80)21 (7.11)
describes the equation of state in the range 0°C I8 5 8"C, Bo = 4°C.
Bennett (1971) included rotation to show that a long shore shear was
generated by the thermal bar density gradients. Further, a transverse
gyre motion supporting the convergence reinforced the thermal bar. This
analysis showed that the time scales for the establishment and propaga-
tion into the center of the lake is of the order of months, distinct from the
behavior in temperate lakes where, as shown above, it is a matter of
hours.
In summary, we have illustrated a new class of convective flows driven
by what has been called differential heating and cooling. The main
features of this flow are that it is three-dimensional and unsteady with the
inertia of the water introducing a large phase lag between the thermal
forcing and the resulting motions. The impact of these motions should
not be underestimated. In the smaller lakes discussed above, these
convective transports will communicate boundary water to the center of
the lake within hours and in large lakes, where rotation is important,
Ookubo et al. (1984) have shown that the boundary thermal inputs lead
to basin scale gyres, which again ensure a greatly enhanced horizontal
transport.

vm. outflow
The outflow from a stratified fluid is influenced by the buoyancy force,
which inhibits differential vertical motion, causing tilting of the isopyc-
nals. When an outlet is suddenly opened, the pressure wave instan-
taneously sets up a radial flow pattern towards the outlet. However, the
radial convergence near the sink quickly distorts the isopycnals, leading
392 Jorg Zmberger and John C. Patterson

to uneven vertical displacements. Buoyancy initiates a set of internal


waves that propagate out, adjusting the isopycnals back to a horizontal
neutral position, and the initial radial flow collapses into a jet-like
structure, flowing horizontally toward the outlet. Such a flow field is
called selective withdrawal, as most of the flow comes from the level of
the sink.
Selective withdrawal has wide application as a management technique
for the supply of water with desired water quality properties. The
technique was used intuitively to scour out poor quality water from the
bottom of the reservoir, long before any theory was developed. It was
not until the early 1960s that Raphael (1962), Koh (1966b), and
Wunderlich and Elder (1969) developed the foundations of the selective
withdrawal process.
The effect of selective withdrawal on the stratification of the reservoir
has received considerably less attention. The thermal structure shown in
Figure 6, measured in Wellington Reservoir in Western Australia, is an
example. Prior to 1976, the lake had a very strong thermocline from
November to June, which caused severe oxygen depletion in the
hypolimnion. In August 1976, the scour policy was first used (see
Imberger and Hebbert (1980)) to rid the reservoir of highly saline water
from the first inflows. The procedure weakened the density structure,
lowering the value of the Lake number so that the thermal structure was
completely changed for the rest of the year. This change influenced the
mixing, which affected the vertical flux of oxygen, and thus the reservoir’s
water quality. While the Wellington Reservoir is an extreme example
because of the influence of salinity on density structure, it displays the
principle well.
Selective withdrawal may thus be used to optimize the water quality in
the reservoir and so obviate the need for mechanical devices such as
pumps or bubblers. Research into selective withdrawal has been inten-
sive, and there is a large body of literature. This discussion is divided into
the following categories:
(1) line sink in a two-layered system,
(2) point sink in a two-layered flow,
(3) line sink in a linear stratification,
(4) point sink in a linear stratification,
(5) selective withdrawal in a rotating stratified fluid.
An empirical procedure for determining withdrawal characteristics from a
Physical Limnology 393

nonlinear stratification is described in Fischer et al. (1979); few fun-


damental results exist.

1. Line Sink in a Two-Layered System

Consider first the simple case of an infinitely deep fluid of density


+
( p Ap) on top of which is a thin, finite layer of water with a density of
p. The case where there is no surface tension between the fluids, the
viscosity is small and a line sink is located a depth H below the interface
was first investigated by Craya (1949). The two layer case is mathemati-
cally identical to that of a single fluid with g' = ( A p / p ) g replacing g
(Imberger, 1980). Since the fluid is infinitely deep and the pressure is
constant along the free surface, the Bernoulli equation can be applied
along the free surface. Craya (1949) postulated a solution with a sharp
cusp extending from the surface to distance h above the sink. This is
distinct from the flow where a small stagnation mount forms over the sink
(Imberger, 1980). The flow at infinity is at rest and, assuming the flow in
the vicinity of the sink is radial and unaffected by the presence of the
cusp, the Bernoulli equation reads

where q is the two-dimensional volume flux at the sink. Craya (1949)


solved this equation for the Froude number F:

such that
F 2 = 8 n 2 ~ 2 (-
1 E),
where E is the ratio h l H . He postulated that the critical Froude number
(the Froude number at which the upper surface is drawn down into the
cusp) was given by
dF2
-- - 0.
dE
This leads to the solution E = $ and

l$ = 3.42. (8.5)
394 Jorg Imberger and John C . Patterson

Gariel (1949) performed experiments with a line sink at the bottom of


a flat bed tank. Modifying his results for the full q as above yields a
critical Froude number
F, = 3.0 (8.6)
Tuck and Vanden Broeck (1984) used conformal mapping and a series
solution technique to determine a cusp solution for the above semi-
infinite fluid domain problem. They obtained a value (for the full q)
F, = 3.55. (8.7)
Hocking (1985) treated the case of a sink above a sloping bottom with an
arbitrary angle. In the limit of a line sink embedded at the apex of the
sloping bottom and the slope going to zero, he found the critical value
F, = 2.88, (8 * 8)
which is smaller but of similar order to the Bernoulli solution (8.6).
Hocking (1988) gave experimental evidence that the critical value was
1.2, much smaller than that predicted by (8.5) or (8.7). Yih (1980),
however, speculated upon the existence of solutions with a cusp to the
problem with a linearized (constant velocity) free surface condition and
hence infinite Froude number. Such solutions are impossible in a fluid of
infinite depth, since the velocity at all points in the fluid must approach
zero at large distances from the sink. This conditipn would mean that the
velocity everywhere on the free surface was zero. If solutions exist in the
finite depth infinite Froude number problem, then solutions must also
exist for finite, but large, Froude numbers. This would mean that for this
geometry, there would be more than one solution with a cusp. Yih (1980)
suggested that the critical withdrawal layer would then be the lowest
value of the Froude number for which a cusp solution existed. Collings
(1985) and Hocking (1988) found solutions, the latter over a complete
range of sink to bottom depth ratios, into the range where the sink rises
above the level of the free surface at infinity (negative ratios).
Vanden Broeck and Keller (1987) found not only the infinite Froude
number solutions but, using a variation on the series method, solutions of
the full range of sink to bottom depth ratios between zero and one, for a
wide range of finite Froude numbers. They found that as the Froude
number was decreased from infinity to a given ratio, cusp-like solutions
were obtained at all values until some critical Froude number was
reached, or the Froude number reached 2. They speculated that where
Physical Limnology 395

4
\
CUSPED SOLUTIONS
\.
1.
1 .
\.
A.

LL"2
CUSPED SOLUTIONS
WITH WAVES -
UNIQUE WAVELESS SPECULATIONONLY
SOLUTIONS

#tEXPERIMENT
HOCKING
(1 988)
0
-l:o -0.5 0.0 0.5 1 .o
(H - h)l H
FIG. 28. Solid line: theoretical results from Vanden Broeck and Keller (1987).
Dot-dash line: theoretical results with a sink (Hocking, 1988). The Froude number F, in
this diagram is based on the depth of water and not the sink depth.

solutions existed down to a Froude number of 2, solutions with waves on


the free surface could also be found for a Froude number less than 2,
although they did not calculate them. They also found the branch of
solutions on which no waves existed and which allowed solutions into the
range of Froude numbers much less than 2. This branch contains
solutions in which the fluid is of infinite depth. These results have been
summarized in the phase diagram in Figure 28.

2. Point Sink in a Two-Layered Fluid

Here, the Bernoulli approach for the semi-infinite domain becomes


(using the same notation)

1
H = h + - - Q2 (8.9)
(4nh2)' 2g'

where Q is the three-dimensional volume flux. Rearranging and finding


the maximum of this expression leads to the critical Froude number

F, = 5.09, (8.10)
396 Jorg Imberger and John C . Patterson
where

(8.11)

This result was first derived by Craya (1949). More recently Wood (1978)
generalized (8.10) to flows in a sector of a sphere and Bryant and Wood
(1976) treated the flow of multiple layers in a sector. Forbes and Hocking
(1988) derived mathematical solutions for the stagnation flow solutions
with no cusps up to a Froude number of 6.4. Experimental verification
was carried out by Lawrence and Imberger (1979) who found a value of
F, = 4.9, which was confirmed by Wood (1978). Harleman et al. (1959)
had found a value of 3.4, whereas the experiments of Jirka and Katavola
(1979) yielded a value of 3.15.
Blake and Ivey (1986a,b) also looked at the case where viscosity is the
controlling force and not inertia to derive the result
h c = 2 . 1 (VgQ
l ) 'I4
9
(8.12)

where h, is the critical draw-down depth. The transition between (8.12)


and (8.10) occurs when the parameter (Ivey and Blake, 1985)
S=(T)Q3g' =2+5.
(8.13)

In both the line and point sinks a diffuse interface will lead to lower
critical Froude numbers (Jirka and Katavola, 1979). This may be
explained by comparing the flow with the upwelling flow in a high
Wedderburn number regime; the three-layer solution of Monismith
(1986) showed a distinct divergence of the interface at the upwind end.
Such a mode 2 response is identical to the interface spreading near the
sink.
For discharges greater than the critical value given by (8.10) or (8.12),
both layers flow toward the sink, such that the interface makes a
well-defined angle at the sink and has a critical point at which the flow in
both layers is controlled. Wood (1978) derived an expression A = Q E / Q H
as a function of the Froude number

where QE is the flow from the epilimnion and QH is the flow from the
Physical Limnology 397

\
\
““I
fly
50

40 -
KINEMATIC APPROACH

KINEMATIC APPROACH
I

30 -

20 -

0
1 01 0 2 °
05 1 2L
10 20 50 100 200 500 1000 2000 5000 10000

F3
FIG.29. The ratio 1 of epilimnion to hypolimnion discharge as a function of the Froude
number for flow towards a point sink. (After Lawrence and Imberger (1979).)

hypolimnion and g’ = (p2- p , ) g / p , . He also presented laboratory find-


ings that verify the concept of the critical point and the flow ratio once
the value of the Froude number was adjusted to compensate for the
non-radial flow at the critical section. Lawrence and Imberger (1979)
carried out further experimental verification of the above theory. Their
results and Woods’ are shown in Figure 29. Murota and Michioku
(1986a) generalized these results to a three-layer fluid.

3. Line Sink in a Linear Stratijication

The horizontal duct is the simplest configuration in which to consider


the initial value problem for a linearly stratified fluid. Initiation of the
sink discharge leads to an instantaneous potential flow solution

(8.14)

where 3 is the streamfunction nondimensionalized with respect to q, and


398 Jorg Zmberger and John C . Patterson

z and x are the vertical and horizontal coordinates nondimensionalized


with respect to H. However, as soon as the fluid particles begin to move,
they experience the buoyancy forces introduced by the stratification.
During these initial stages, inertia and viscous forces will be negligible
and the flow represents a balance between the unsteady vorticity term
and the baroclinic generation of vorticity as expressed by the streamfunc-
tion equation:
V21Vn + lVxx = 09 (8.15)
where 3 = V / q , z = z / H , x = x/H, t = tN, N is the buoyancy frequency
of the stratification, and V2 is the Laplacian operator.
Imberger (1970) used transform techniques to verify that close to the
sink, the solution to the duct problem was identical to the solution in the
unbounded fluid domain solved earlier by Koh (1966a). These authors
were able to show that for large time the horizontal velocity is

where = sin 8, 8 = arctan(z/x), r = (x’ + z ~ ) ” ~and , Jo is the Bessel


function of zeroth order. In the limit t + (x ~ and z fixed), this solution
reduces to
u(x, z , t ) - 4 2 ) ) (8.17)
which represents a singular jet flow at the level of the sink. The internal
waves responsible for changing the initial uniform flow in the duct into
the line jet must move at finite speed and hence steady state can never be
achieved in the model problem described by (8.15).
It was not until the work described in Kao et al. (1974) and Pao and
Kao (1974) that the role of the internal waves was fully understood.
These authors simplified (8.15) by omitting the x derivatives in the
Laplacian operator and showed that the double Fourier series describing
both the horizontal and vertical wave number space could be simplified to
a single sum of columnar disturbances.

@=
1
2>
2 “ 2
(4- + n = l -
nn
sin 2nnz H(nnx + t). (8.18)

McEwan and Baines (1974) independently defined these columnar


disturbances, but named them shear waves, in connection with a different
problem. They derived, by a modified method of steepest descent, the
behavior of the fluid in the vicinity of wave fronts. Their analysis showed
Physical Limnology 399

that instead of remaining compact, as suggested by the step function in


(8.18), the shear waves are analogous to expansion waves in compressible
flow, and a wave travelling with a speed c,, = H N / n n has a frontal width
w,,= ~ ( ~ t ) ” ~ / n .
Physically, the initiation of the sink discharge thus leads to a potential
flow which, in turn, may be decomposed into a complete set of
progressive internal waves, containing the whole spectrum of vertical and
horizontal wave numbers. Once established this initial wave field
propagates away from the sink leaving a steady sink flow in its wake. As
Mahony and Pritchard (1977) point out the sole purpose of these waves is
to adjust the isopycnals so that they can return to a horizontal position at
steady state. McEwan and Baines (1974) proved that it is possible to sum
over the horizontal wave number space leaving an infinite series of shear
waves.
It may now be shown (Clarke, pesonal communication) that the long
wave limit of (8.15) is the Airy Equation (Cohen, 1979) and so shear
waves may be thought of as solutions to the Airy Equation. Clarke has
also recently shown that the inclusion of the nonlinear convective
acceleration term leads to the Korteweg-de Vries equation (Benney,
1966; Djordjevic and Redekopp, 1978) so that the nonlinear aspects of
shear waves can be thought of as analogous to an undular bore (Smyth
and Holloway, 1988).
The nonlinear aspects of a shear wave generation, distinct from the
propagating properties, were investigated experimentally by Monismith et
al. (1988). As the inertia of the flow increased, the observed shear wave
amplitude decreased relative to their theoretical values. When the
outflow was repeatedly pulsed, the strength of the average flow towards
the sink indicated a withdrawal layer thicker than that which would have
been found given either the average flow rate or the maximum
instantaneous flow rate.
Each shear wave has a vertical wave length A, = 2H/n and a wave
celerity C, = AnN/2n, so that the approach to the 6 function velocity
distribution given by (8.17) is brought about by the passage of successive
shear waves, each with a progressively smaller vertical wave length.
Steady state is achieved either by the dissipation of these waves or by
advection sweeping the waves back out through the sink.
First, consider the action of viscosity, which will dissipate the nth wave
in a time of O(Ai/v). However, in this time the wave has moved the
400 Jorg Imberger and John C. Patterson

distance A;N/2nv so that the non-dimensional thickness 6 / x becomes

-6 --_
An-
- O(Gr;1'6), (8.19)
x 2 x
where Gr, = N2x4/v2. Thus, a viscosity dominated layer has an x'13
upstream growth. This layer was first derived theoretically and verified
experimentally by Koh (1966a). The velocity profile is bell shaped with a
weak return flow immediately above and below the main layer.
Second, as a shear adds to its predecessor, the central velocity growth
is of order 2nxlNt. Advection will dominate once the induced velocity
O(v/An) is equal to the nth wave celerity O(AnN/2n). Successive waves
of order greater than n will no longer be able to propagate upstream.
This condition leads to a layer thickness
6
- = O(F;I2), (8.20)
X

where F, = q / 2 N x 2 and q is the total flow from both sides. Hence, in the
inertial case, the layer is of constant thickness converging only in the
potential flow region near the sink.
Combining (8.19) and (8.20) shows that the withdrawal layer changes
from one dominated by viscous forces to an inertially dominated layer at
a distance x , given by:

(8.21)

where RL = FL Grz3 and L is a reference length replacing x . The length x ,


was first recognized by Imberger (1972), and Silvester (1979) confirmed
its importance with a series of laboratory experiments.
The viscous flow given by (8.19) will no longer form the correct
upstream flow when the experimental tank is shorter than x , . Debler
(1959) described experiments in a short tank with an adjoining large
upstream reservoir. In such circumstances, pU2 is constant upstream, and
the motion is governed by the Euler equation together with the constraint
DplDt = 0. As demonstrated by Dubreil-Jacotin (1934), this leads to the
modified streamfunction equation

VqJ' = gz
- 7 +7 ,
d p dH
(8.22)
PodW d v
where q ~ =
' (p/po)1'2qJand A is the total upstream head. The first term
Physical Limnology 40 1

on the right side of (8.22) represents the vorticity generated by the


buoyancy forces, while d H / d v ’ is the vorticity inherited from the
upstream boundary condition. Yih (1965) showed that if pU2 = constant
and the density varies linearly with z , then the right side of (8.22) is
constant and a solution may be found by separation of variables for
F = q / 2 N H 2> l/n. This solution consists of uniform flow upstream
necking down to a withdrawal layer near the sink. Above and below this
layer are weakly rotating eddies, which extend to infinity at F = l / n , thus
violating the upstream boundary conditions. This result was also derived
by Trustrum (1964). Kao (1970) constructed a numerical solution for
F < 1/n, along the lines of the method used by Huber (1960). Kao (1970)
showed that the layer thickness was given by (8.20), with a numerical
constant l h instead of the order sign. This value agrees with the
experimental results of Debler (1959).
In all experimental and prototype applications of the theory, some type
of rear wall exists that blocks the horizontal velocities. Imberger and
Fandry (1975) investigated the influence of such blocking by solving the
flow in a vertical duct. They demonstrated that internal waves were
indeed initiated but the vertical duct prevented their separation into
vertical shear waves. The internal wave group merely moved vertically
through a uniform flow, leaving behind the flow into the sink. This model
showed that the horizontal velocity varied linearly from x = 0 to x = L / 2 .
Imberger et al. (1976) solved the flow in a finite tank with a free
surface. The vertical constraints reintroduced shear waves that traversed
the tank, reflecting at the vertical boundaries. Their solution highlighted
a number of points:
The horizontal jet discharge varied linearly from the sink to the
end wall.
For the parameter R L < 1, the withdrawal layer thickness was
given by
6
- = 5.5 Gr,1’6 pr-1’6, (8.23)
L
with 64% of the layer lying above the sink center line and where Pr
is the Prandtl number.
For RL > 1, the flow was completely inertial and the withdrawal
layer thickness was given by
s
- = 4.OFZn. (8.24)
L
402 Jorg Imberger and John C . Patterson

(d) The withdrawal layer flow may be assumed to be steady through-


out the tank when it is steady at the end of the tank. This led to a
set-up time t = O(N-’Grg6) for RL < 1 and t = O(N-1FL1/2)for
R,> 1.
These conclusions, including the numerical factors in (8.23) and (8.24),
have been verified experimentally by Silvester (1979). Further, data given
by Bohan and Grace (1973) for the inertial flow regime may be
interpreted via the above formula, (8.24) yielding a numerical factor
of 3.4.
In the above, it was implicitly assumed that the Prandtl number was of
order unity. In many applications, especially where the water is stratified
with salt, this is not the case. Imberger et al. (1976) showed that the flow
structure was much more complicated for fluids with a large Prandtl or
Schmidt number. Basically, two complications arise. First, there are two
critical horizontal length scales x , , one where the viscous buoyancy
diffusive layer changes to a nondiffusive layer and one where the
influence of inertia becomes important. Second, the layer of thickness
order (Gr-’I6) is no longer stable but undergoes a secondary and tertiary
collapse. This leads to a hierarchy of layer types depending on the
relative magnitude of RL compared to the Prandtl (Schmidt) number Pr.
One of these layers has a thickness 6 / L = O(Gr-1’6R~s),experimentally
found by Mahony and Pritchard (1977), and one that is basically the
nondiffusive layer documented by Gelhar and Mascolo (1966) and
Walesh and Monkmeyer (1973). Silvester (1979) has experimentally
verified the existence of the flow hierarchy, but there remains a difficulty
in the regime Pr-’16 < R L < Pr-’13, where the potential energy of the flow
is zero over certain parts of the layer (see Imberger and Fischer (1970))
and where Mahony and Pritchard (1977) have observed hysterisis effects.

4. Point Sink in a Linear Stratification

The axisymmetric withdrawal layer was first investigated by Koh (1966a)


in the limit of very small discharge so that the flow was governed by a
viscous buoyancy balance. Lawrence and Imberger (1979) demonstrated
that this steady flow was established by the passage of a series of shear
waves originating from the sink at the initiation of the outflow. However,
in the axisymmetric case, these waves are cylindrical with amplitudes
depending inversely on the radius of propagation. These authors postu-
Physical Limnology 403

lated, since the layer thickness was independent of the discharge, that the
thickness should be the same as in the two-dimensional case. This is
borne out by the data from Koh (1966a), who found
6
- = 5.8 Gr-116 pr-1/6 (8.25)
L
for the case where inertia was negligible. Spigel and Farrant (1984)
investigated the inertia-buoyancy limit and verified that the layer
thickness was given by the scale presented by Imberger (1980):
113
6=C@ 9 (8.26)

where C1= 1.58 (Bohan and Grace, 1973), 1.32 (Lawrence and Imber-
ger, 1979), 1.6 (Spigel and Farrant, 1984), and 1.32 (Ivey and Blake,
1985). Ivey and Blake (1985) also showed there was a layer thickness
6 = 4 . 2 (vQ
3)
(8.27)

that described selective withdrawal when inertia was small but species
convection was important.
The transition between these various layers is more complicated than
in the two-dimensional case. Ivey and Blake (1985) found a transition
parameter
(8.28)

but this parameter is now independent of distance from the sink so that
the inertia-free convective regime thickness (8.27) is also no longer
dependent on distance from the sink. A double criterion thus arises
depending on whether the convective layer can exist at all. These are
A: S > 3, S > (Gr Pr-2)1130 Inertial layer (8.26);
B: S > 3, S < (Gr P Y -~ )” ~ ’ Diffusive layer (8.25);
C: S < 3, S > (Gr Pr-2)1118 Convective layer (8.27);
D: S C 3, S < (Gr Pr-2)1118 Diffusive layer (8.25);
The difference here is that in Cases C and D, the inertial layer is thinner
than the convective layer and thus cannot exist. The first criterion on S is
a measure of the relative size of the inertial and convective layers, while
the second criterion on S differentiates the inertia and diffusive layers in
Cases A and B and the convective and diffusive layers in Cases C and D.
404 Jorg Imberger and John C . Patterson

5. Selective Withdrawal in a Rotating Stratijied Fluid

A homogeneous rotating fluid behaves analogously to a stratified


rotating fluid (Veronis, 1967), so it is not surprising that selective
withdrawal has its counterpart in a rotating fluid. Long (1956) obtained a
solution for steady flow of a rotating inviscid fluid into a point sink
located at the axis of the rotation to an infinitely tall cylinder. Using the
full nonlinear equations (Yih, 1965), he found necking near the sink with
an annular eddy around the axisymmetric layer. This eddy elongated to
infinity as the Rossby number

(8.29)

(where Q is the discharge, b the radius of the cylinder, and f the Coriolis
parameter) approached 0.261. Shih and Pao (1971) carried out experi-
ments and found that the withdrawal layer formed with a radius

(8.30)

whenever Ro < 0.261. Motivated by these experiments, Pao and Shih


(1973) constructed a slip line solution and found excellent agreement with
the prediction from (8.30). The viscous dominated solution was also
derived by Pao and Kao (1969) and they found in this case that the layer
thickness was given by
113
&=4.1(7) , (8.31)

the numerical fctor being obtained both theoretically and experimentally.


The transition between the two does not appear to have been studied,
but by analogy with a purely stratified case, the transition should depend
on the parameter

R --.Q (8.32)
-YL
For Rf > 1, (8.30) applies; for Rf < 1, (8.31) is relevant.
The above formulae predict a vertical withdrawal layer whereas in the
stratified case, the withdrawal layer is horizontal. Investigations are
underway (McDonald, personal communication) in an effort to reveal the
transition of the flow as the parameter f /N is varied from 0 to infinity.
Physical Limnology 405

Monismith and Maxworthy (1988) performed experiments on the estab-


lishment of a selective withdrawal layer in a rectangular channel under
conditions of strong stratification and weak rotation ( f l N = 0.8- 0.25).
Their observations show that the initiation of the outflow resulted in a
Kelvin shear wave propagating cyclonically around the perimeter of the
tank, leaving in its wake an anti-cyclonic withdrawal layer. The thickness
of the withdrawal layer was, however, not influenced by the rotation and
was given by (8.26).

In summary, a great deal is known about selective withdrawal and most


situations can be predicted. There are three exceptions, however. First,
arbitrary stratification leads to premature withdrawal because of the
higher mode response, but no predictive procedure exists. Second, the
influence of sills or contractions has received only preliminary treatment
(Kao, 1976) and much work remains to be done. In particular, transitions
that occur through a contraction where the withdrawal layer changes
from a subcritical condition to supercritical and back to subcritical with
an associated hydraulic jump would appear to be extremely important to
mixing in lakes and the ocean. Third, the transition flow from a fully
rotating fluid to a fully stratified fluid has not been studied in detail.

Ix. Inflow

Reservoirs and lakes are supplied from rivers; as the river water meets
the relatively stagnant reservoir water, it usually encounters water of a
slightly different temperature, salinity or turbidity. These variables lead
to an inflow that is either lighter or heavier than the surface water of the
lake. When the inflowing water is very much heavier (Hebbert et al.,
1979; Ford et al., 1980) the river water enters the reservoir, and plunges
to the bottom of the lake. Alternatively, examples of where the inflow
water was not sufficiently heavy to take it all the way to the bottom are
given by Howard (1953), Wunderlich and Elder (1969), Ford (1978),
Ford and Johnson (1981), Fischer and Smith (1983), and Imberger
(1985). In all these cases, the inflowing water plunged beneath the lake
water at the entrance then flowed down the drowned river valley,
entraining lake water as it moved downstream. At some depth, the total
entrainment made the underflow neutrally buoyant relative to the
immediately adjacent water, and the underflow became an intrusion
406 Jorg Zmberger and John C . Patterson

FIG. 30. Schematic of inflow possibilities: overflow, plunge-point, intrusion, and


underflow.

(Figure 30). Last, there is the case where the inflowing water is lighter
than the resident surface water and the river overflows the lake as a
buoyant surface flow (Figure 30). The dynamics of such an overflow are
similar to the situation where a source of fresh water enters an estuary or
coastal environment (Chen, 1980; Ookubo and Muramota, 1981; Chao,
1988; Luketina and Imberger, 1987, 1988).
The surface buoyant jet occurs only rarely in a lake in tropical and
temperate regions but is normal in high latitude cold lakes, where such
inflows induce thermal bars (see Section VII). The dynamics of these
flows is a subject in itself and has been extensively treated by Chen
(1980), Sargent and Jirka (1982), Fischer and Smith (1983), O’Donnell
and Garvine (1983), Chu and Jirka (1985), Arita et al. (1986), and
Luketina and Imberger (1987, 1988), with an excellent review of the
frontal region being given by Simpson (1982). The reader is referred to
these articles for further information on the overflow case.
Consider the case where the density of the inflowing water is somewhat
heavier than the resident surface water in the lake and the river water
will plunge down the river bed until it reaches neutral depth. The
Physical Limnology 407

propagation down the inclining plane of the leading head of the gravity
current has been investigated by Stacey and Bowen (1988) and Wilkinson
and Wood (1972). However, in most cases, the inflowing discharge is
sufficiently steady to allow the underflowing river to be analyzed as a
steady or quasi-steady gradually varying flow.
In this case, a balance is struck in the underflow such that bottom
friction, interfacial entrainment and inertia retard the flow, and the
downslope component of gravity accelerates it. This momentum balance
is influenced by the slope and flatness of the river bed and the rotation of
the earth. A discussion of the latter influence is omitted here.
As mentioned above, most authors have assumed the underflow
discharge to be steady, and if it is further assumed that the underflow is
confined laterally by the drowned river valley, a gradually varying flow
analysis yields satisfactory results (Ellison and Turner, 1959; Savage and
Brimberg, 1975; Hebbert et al., 1979; Akiyama and Stefan, 1984). The
main conclusion reached in these studies (see Fischer et al. (1979)) is that
the flow quickly adjusts, so that as it propagates downstream along a
particular bed slope, the Froude number of the underflow becomes a
constant equal to the normal flow Froude number F,, where the Froude
number is given by

where u is the velocity of the downflow, g' is the modified acceleration


due to gravity between the lake and underflowing water, and d is the
hydraulic depth of the underflow.
For the case where the density difference is derived from particulate
matter (turbidity current), the fallout or accumulation of particles must
be accounted for. However, provided the rate of change of density due to
deposition or erosion is slow compared to the time it takes to travel a
number of depths of the underflow, it is possible to retain the gradually
varying flow assumption, so that (9.1) is still valid even if only locally
(Akiyama and Stefan, 1985). Fischer et al. (1979) solved the vertically
integrated momentum and mass equation to derive an expression for the
normal Froude number:
408 Jorg Zmberger and John C. Patterson

where T is the top width of the underflow, E is the entrainment


coefficient, P is the perimeter of the underflow, h is the underflow depth,
h, is the depth of the underflow centroid below the interface, a is the area
of the underflow, 9 the valley slope, and CD is the channel bottom drag
coefficient.
The entrainment coefficient may be determined by applying a model
similar to the one used for the surface layer (see Section IV), but as
discussed in Fischer et al. (1979), the coefficients of the model may be
different due to the very different velocity profiles within the underflow-
ing current compared to the velocity distribution in the wind driven
surface layer. The entrainment at the interface is energized by the shear
there and by the turbulence imported from that generated at the river
bed. For mild slopes (9 = Hebbert et al. (1979) have shown that
shear production was negligible and that E had a value of 1.9 x
The location of the plunge point can now be found by neglecting the
local entrainment. This is done by substituting the entrance flow and
density difference into (9.1) and equating this to the normal Froude
number evaluated from (9.2). This will yield the first estimate of the
plunging depth h l , which may be improved by accounting for the
non-hydrostatic pressure distribution near the plunge point (Wilkinson,
1972; Hebbert et al., 1979; Jain, 1982; Akiyama and Stefan, 1984, 1985,
1987). This leads to a relationship

which implies an energy loss at the plunge point.


The three equations (9.1 to 9.3) can now be solved together for d and
H with only the assumption of no entrainment at the plunge point.
Comparison with field data is excellent (Hebbert et al., 1979).
When the slope becomes steeper (F, larger) or the inflowing momen-
tum is increased, the energy loss implied by (9.3) can no longer be
neglected and the associated entrainment at the plunge point will
invalidate the above analysis and the theory must be modified as
suggested by Britter and Simpson (1978) (see also Luketina and Imberger
(1987) and Jirka and Arita (1987)). The analysis for the inclusion for such
entrainment has been carried out by Akiyama and Stefan (1987), but no
data exist to fix the amount of entrainment except for the buoyant
overflow documented in Luketina and Imberger (1987, 1988). There are
some numerical simulations available (Farrell and Stefan, 1986; Buchak
Physical Limnology 409

and Edinger 1984). However, these all use closure schemes so that the
estimates of entrainment would be overestimated (Franke et al., 1987).
For a divergent river valley with a relatively flat bottom, the flow may
separate from the side walls. Johnson et al. (1987a,b) observed six
different flow regimes depending on the entrance aspect ratio, the Froude
number, and the angle of divergence. These are schematically re-
produced in Figure 31. In general, the higher the Froude number, the
closer the flow is to a simple, entering jet. Once separated from the side
wall, the inflow behaves as a free jet and Johnson ef al. (1988) found that
the entrainment was much higher for such separated flows and could be
estimated from simple jet analysis (Fischer et al., 1979; Rodi, 1982).
The case of inflow into a lake with a flat bed morphology is similar to
the large divergence case and is characterized by a central momentum jet
followed by a radially spreading buoyant plume, originating from a
virtual origin. The virtual origin occurs at a distance from the inflow of
order M3/4B-112(Fischer et al., 1979; Luketina and Imberger, 1987),
where M is the entrance momentum flux and B is the corresponding
buoyancy flux. Luketina and Imberger (1987, 1988) presented an analysis
and a comparison with field data of a buoyant overflow, whereas
Hauenstein and Dracos (1984) presented a similar analysis and a
comparison with laboratory data for the case when the inflow is denser

III-IMMhA
SMALL DIFFUSER ANGLE LARGE
S T R O N P BUOYANCY WEAK

SURFACE
PLUNGELINE (REGION)

A1 A2 A3 B C E

STREAMLINES

FIG.31. Principal regimes of negatively buoyant flow into horizontal diverging channel.
(After Johnson er al. (1987a).)
410 Jorg Zmberger and John C. Patterson

than the lake water. Recently, Tsihrintzis and Alavian (1987) developed
a further integral model similar to that of Hauenstein and Dracos (1984)
for three-dimensional flow down an inclined plane.
Once the inflow reaches a depth of neutral buoyancy, the inflowing
water leaves the river valley and intrudes horizontally into a stratified
lake. Imberger (1985) presents field data from the Wellington Reservoir
that illustrate this lift-off process. The dynamics of the intrusion appear to
have only been thoroughly investigated for two-dimensional flow. In this
case, an intrusion entering a linearly stratified ambient water body will
have a length and width as summarized by Imberger et al. (1976) with the
propagation length being given by

i
0.44 LRli2t’, t’ 5 R, (9.4)
I = 0.57 LR2/3t’5i6, R -=t’ <PI?, (9.5)
CLR 3/4f’3i4, P? < t‘ < R-’, (9.6)
where the coefficients in (9.4) to (9.6) were determined from experimen-
tal data by Manins (1976) and Maxworthy (1980), but where the constant
C in (9.6) is still unknown. In these formulations, the parameters are
defined as
R = F Gr’I3, (9.7)

N2L4
Gr=- (9.9)
Y2 ’

(9.10)

and I is the length of the intrusion at a time t after initiation of the flow.
As already discussed for intrusions formed by differential deepening,
(9.4) to (9.6) are not valid when the intrusion approaches the rear wall
(Fischer et al., 1979).
Darden et al. (1975) showed that the accompanying velocity structure
due to such an intrusion consists of multiple reversing jets above and
below the entering intrusion, and Maxworthy (1980) demonstrated that
simple intrusions can degenerate into solitons if the volume flux to the
intrusion is terminated.
The second main case in a lake is where the downflow encounters a
Physical Limnology 411

sudden change of density at a density interface. If the water below the


density interface is heavier than that of the downflowing water, an
intrusion will form along the interface. If the interface is very thin
compared to intrusion thickness, the intrusion behaves as two separate
gravity currents, the depth of the intrusion relative to the interface
adjusting so that the speeds match (Holyer and Huppert, 1980; Britter
and Simpson, 1981). Faust (1981) gives the speed of propagation V as

where d , is the depth of the upper layer, d2 is the depth of the lower
layer, p1 is the density of the upper layer, p2 is the density of the lower
layer, hl is the depth of the upper fluid above the intrusion, h2 is the
depth of the lower layer below the intrusion, and Ap = p2 - p,. Faust
(1981) generalized the same analysis and showed that a similar expression
could be derived for an intrusion into a three-layer fluid. However, in this
case, he found that internal waves generated at the interfaces could lead
to a slow-down of the intrusion.
The above intrusion dynamics have only considered laminar flow.
However, in lakes, the downflow is strongly turbulent (Imberger, 1985)
and it is important to assess the collapse of this turbulence as the
intrusion enters the lake. Lin and Pao (1979) and later Hopfinger (1987)
have reviewed this subject and concluded that the collapse takes place in
times of O(N-') and that the vertical motions decay most quickly,
leaving, in the absence of additional shear, two-dimensional turbulence
with a vorticity vector orientated predominantly vertically. The boundary
mixing experiments of Browand ef al. (1987) suggest that the decay is
extremely rapid and that we would not expect to find turbulence within
the intrusion at a horizontal distance from the lift-off point much greater
than the thickness of the intrusion. This has been confirmed by Roberts
and Matthews (1987) who injected a horizontally neutral buoyant
momentum source from a round nozzle into a linearly stratified fluid and
found that the mean shape of the jets spread horizontally at a distance of
M"4N-1",where M is the momentum flux of the jet and N is the
buoyancy frequency of the ambient water. Although these authors did
not measure the turbulent flux directly, they inferred that the largest
turbulent eddies became anisotropic at the same distance.
Now in the case of an underflow, the Froude number based on the
density difference between the underflow and the neighboring ambient
412 Jorg Imberger and John C . Patterson

water becomes infinity since g‘ tends to zero at the point of neutral


buoyancy. However, the inertia of the underflow would carry the fluid
through the lift-off into the intrusion, and a finite sized intrusion based on
an inertia buoyancy balance would form with a thickness (Manins, 1976)
112
6 = 4);( (9.12)

for the two-dimensional flow case. Assuming that the collapse starts at a
distance order (MN-2)”3, which is the equivalent result for the two-
dimensional flow, the turbulence would commence the decay at a
distance O(QN-l”) or one intrusional layer thickness into the lake.
Given these two pieces of evidence, one would expect not to find
turbulence in an intrusion. However, recently Imberger (1987) has
shown, by carrying out temperature microstructure measurements, that
the shear from the intrusions together with background shear from
internal wave activity can sustain a low level of turbulence. In a further
analysis of the same data, Imberger (1988) has shown that the turbulence
was at the point of decay, with

Fr, = (5) 112


= 15ll2= 3.87. (9.13)

Thus, even though the gradient Richardson number associated with


intrusion was of the order of 11, the turbulence remained viable,
although at the point of fossilization (Rohr et al., 1988).
In summary, we have seen that the inflow dynamics may be con-
veniently partitioned into three separate problems: the underflow, the
plunge point, and the intrusion dynamics. The inflow of the heavier water
mass into a drowned river valley appears to be adequately described by
existing gradually varying flow theories, although the entrainment over
steep slopes where shear production is the dominant source of energy for
entrainment has not been verified in the field. Equally, under these
conditions, we may expect severe entrainment at the plunge point, which
so far defies parameterization. The interaction of the inflow with a
resident internal wave field (Fischer and Smith, 1983) or with a surface
mixed layer deepening (Imberger , 1985) can introduce enhanced mixing
and must be treated on a case by case basis for each lake.
The most important problem remaining concerns the mixing associated
with inflows, both directly in the intrusion and also, as discussed in the
next section, in the lake as a whole.
Physical Limnology 413

X. Mixing below the Surface Layer

Mixing below the surface layer is generally patchy and sporadic, with
individual events occupying no more than tens of cubic meters and
remaining at full intensity for no more than a few minutes (Imberger,
1985). The two central questions remain: What is the efficiency of energy
conversion at turbulent sites and what is the patch statistical distribution
(Imberger and Hamblin, 1982)?
No direct measurements of the local value of the vertical diffusion
coefficient K , exist, but many investigators have taken advantage of the
closed nature of the lake basin and obtained estimates of the vertical
diffusion coefficient averaged over the whole basin (Kullenburg et al.,
1974; Jassby and Powell, 1975; Imboden and Emerson, 1977; Imboden et
al., 1977; Robarts and Ward, 1978; Weiss et al., 1979; Lewis and Perkin,
1982; Imboden et al., 1983; Imboden and Joller, 1984; Sanderson et al.,
1986; Colman and Armstrong, 1987; Lake Biwa Research Institute, 1987;
Wuest, 1987; Wuest et a/., 1988). The data confirmed (Imberger and
Hamblin, 1982) that the lake-wide averaged vertical exchange coefficient
fluctuates between 2 x m2 s-l (near molecular) during periods of
very strong stratification and weak winds (Wuest, 1987) to values as high
as m2s-l during periods of very weak stratification in tropical lakes
(Lewis and Perkin, 1982). It has also been shown (Ward, 1977;
Straskraba, 1980; Wuest, 1987) that the vertical diffusivity depends on
the buoyancy frequency, the area of the lake and the depth of the lake.
Wuest (1987) attempted a correlation of the rate of mixing with the
Wedderburn number (2.5) and the buoyancy frequency, writing the
vertical diffusion coefficient as

(10.1)

He found most data sets required a value of cy in the range of


0.3 < LY < 0.9. This may be compared with oceanic variations, summar-
ized by Gargett and Holloway (1984), where 0.25<a<0.5. The
dependence on the Wedderburn number is shown in Figure 32, re-
produced from Wuest (1987), and indicates that if W ranges from
between 200 and 0, the vertical diffusivities range from around
10-6m2s-' to around 10-*m2s-'. In computing the value of the
Wedderburn number W, Wuest (1987) used the depth of the epilimnion,
but does not give details of the definition used. However, judging from
414 Jorg Zmberger and John C. Patterson

106 10-4 10.2

K, (rn21s)

FIG.32. A comparison of K , in Sempachersee for four different periods. The value of


K , was calculated using traces and are typical of the three regimes discussed in the text:
0.5 < W < 200, 0.1 < W < 0.5, and W < 0.1. (After Wuest (1987).)

his profile shapes, this would have been close to the depth of the seasonal
thermocline, which would make W close to LN. His W is really an
indicator of the stability of the seasonal thermocline and not the surface
layer.
We saw in Section I1 that in the majority of cases a lake is adequately
approximated by a three-layer stratification, so there are two numbers, W
and LN , which determine its response to a wind force. Small W and large
LN lead to a strong surface layer response, and small LN causes the main
thermocline to tilt with the associated stronger mixing in the hypolim-
nion. Three examples provide an illustration. In Figure 33, data from the
Canning Reservoir show a case where W =0.1 and LN=288. The
dissipation levels are close to zero everywhere, except in the top two
meters, where a very weakly stratified surface layer resided. During the
time of this profile, nearly 200 microstructure casts were taken over a
period of 10 days and only one profile showed the slightest sign of
turbulence in the hypolimnion. By contrast, Figure 17f depicts a
microstructure profile at W = 2.1 and LN = 0.24, whereas in Figure 24c
the conditions were such that W = 4.5 and LN = 0.44; in both cases the
whole profile had become turbulent with dissipation values in the
hypolimnion reaching m2s - ~ . These data are presently being
analyzed more fully, but this preliminary analysis, together with the
evidence from Wuest (1987), strongly suggests that the response of the
Physical Limnology 415

TEMPERATURE F) fldr ( a m ) DISSIPATON E (mas31

FIG.33. Microstructure taken in the Canning Reservoir during a period of large Lake
number L , = 288. (a) Temperature. (b) Temperature gradient. (c) Dissipation as calculated
from the Wigner-Ville technique (sticks) and the Batchelor spectra fitting technique (bar
chart). Note that there is no turbulence below the top 2 m.

seasonal thermocline, which determines the deep mixing in a lake, is in


general determined by the value of the Lake number LN.
The variability within the hypolimnion must depend on some empirical
relationship similar to that given by (10.1). Weinstock (1984) has
determined the following theoretical formulation:
&
K, = a (10.2)
N2+ k;u2 '
where E is the dissipation of turbulent kinetic energy, N is the buoyancy
frequency, ko is the wave number of the energy bearing eddies, u is the
root mean square velocity of the motion in the turbulent patch, and a is a
constant between 0.2 and 0.8 (Imberger and Hamblin, 1982). It is clear,
however, that the dependence on N is too strong in (10.2) and is not
supported by the empirical measurements of the average diffusion
coefficient (10.1). However, it would be interesting to investigate whether
the averaging process could decrease the dependency from N-' to N-'.
416 Jorg Imberger and John C . Patterson

The following overall scenario is appealing: as the wind, inflow, or


artificial mixing device (e.g., an aerator) start up, they induce seiching of
the thermocline (mode 1 response in a three-layer system). This seiching
provides a background shear within which internal waves generated by
differential deepening and by leakage through the base of the surface
layer increase the frequency and intensity of turbulent patches. This leads
to an increase in the buoyancy flux and turbulent kinetic energy
dissipation. This is a much simpler picture than in the ocean where the
background shear arises from a host of motions.
The Lake number can be extended easily to account for these three
major disturbances to read

(10.3)

where Qo is the air flow rate of the aerator, Qi is the river inflow rate at
the point where the intrusion forms, U is the inflow velocity of the
intrusion, and zi is the height of the intruding intrusion. All other terms
are defined in Sections I1 and XII. The additional terms are the
additional disturbing moments offered by the bubble flow pog2nQ$'3~g3A0
(Section XII) and the river momentum flux pOUQiacting at a moment of
(zg- zi) (Section XI). Whenever LN is small, the seasonal thermocline
may be expected to respond more actively, thus leading to higher
transport rates.
The origins of the mechanisms leading to internal mixing are numer-
ous, consisting of a range of instabilities, wave-wave, wave-shear,
Kelvin-Helmholtz billows, Holmboe instabilities, critical layer absorp-
tion, double diffusion, bores, and hydraulic jumps. These mechanisms
have recently been thoroughly reviewed and the reader is referred to
Hopfinger (1987) and Thorpe (1987). Mixing in the metalimnion is also
reviewed by Gregg (1987). For a review of double diffusion, the reader is
referred to Turner (1985).
The nature of the turbulence in a lake can be determined from the
turbulent Froude versus Reynolds number diagrams (Gibson, 1980, 1982,
1986, 1987a,b, 1988; Imberger and Boashash, 1986; Imberger, 1988). The
Froude number is defined

(10.4)
Physical Limnology 417

and the Reynolds number is given by

(10.5)

where E is the dissipation of the turbulent kinetic energy, g' is the


acceleration due to gravity derived from the density anomaly within the
turbulence, and L, is the displacement scale (Imberger and Boashash,
1986). Turbulence close to the line
( T )= 3.87
Fr, = EL,
112

(10.6)

is at the point of fossilization (Itsweire et al., 1986; Imberger, 1988),


whereas active turbulence lies in the region Fr, > 1 and Re > 1. In Figure
34, we reproduce data derived from turbulence measurements from 10
different physical mechanisms operating in lakes. The mixed layer
turbulence is active; turbulence in the hypolimnion is at the point of
decay. It is necessary, and possible, to design a field experiment aimed at
correlating the location of the data points in the Froude versus Reynolds
number phase diagrams with the values of W and LN.Given that the
mixing efficiency (the flux Richardson number) can be correlated to the
location of the data in Figure 34, this would lead to an operational model
for the prediction of vertical transport in a lake (Imberger, 1988).
So far we have discussed only mixing internal to the hypolimnion.
There is evidence in the ocean (hey, 1987; Garrett and Gilbert, 1988;
Gregg and Sanford, 1988) that the energy input due to tidal motion
exceeds that appearing in the water column at depth, suggesting that
dissipation in the benthic boundary layer is a major sink for this energy.
Lueck and Osborn (1985) presented evidence that close to the bottom in
a submarine canyon, dissipation values increased considerably. h e y and
Boyce (1982) showed that a benthic boundary layer existed at the bottom
of Lake Erie.
In Figure 35, we show the development of turbulence in the water
column of the Harding Reservoir in response to a surface wind stress on
27 November 1988. Early in the morning the meteorological conditions
were warm and calm. At 1000 hours a wind with a speed ranging between
2 and 3ms-' commenced and continued until 1400hours, leading to a
Wedderburn number of 0.05 and a Lake number of 2.2. The sequence of
profiles (all extending from the bottom to the lake surface) shows the
establishment of turbulence both with the hypolimnion and in the benthic
418 Jorg Imberger and John C . Patterson
Physical Limnology 419

boundary layer with dissipation values as high as 6 x m2s - ~in the


bottom 1 m of the profile (Figure 35f). This data set is presently being
analyzed but was included as it offers the first conclusive evidence of the
development of turbulence in the benthic boundary layer of a lake.
It is tempting to carry out a budget of the turbulent kinetic energy. For
instance, the data shown in Figure 33 indicate dissipation in the surface
layer (depth of 2 m) that almost exactly balances the energy input CLu:.
However, by comparison, integration of the total dissipation per m2 of
the lake’s surface for the profile shown in Figures 17 and 24 greatly
exceeds that introduced at the surface. Sufficient measurements are
required to account for the very strong variability (see Section VI) of
wind stress over the lake surface.
Boundary mixing can be sustained by two major mechanisms. First,
internal wave seiching will cause a velocity shear at the boundary and set
up a turbulent boundary layer over the rough lake bottom (see Imberger
and Hamblin (1982) for a review). Numerous experiments (Ivey and
Corcos, 1982; Thorpe, 1982; Phillips et al., 1986; Browand et al., 1987;
Ivey, 1987) have modeled this mechanism with a grid stirring the water
adjacent to a vertical or sloping wall in a stratified tank. These
experiments all produce a turbulent boundary layer close to the grid,
which then initiates intrusions that flow horizontally into the tank. The
mean motion induced by these intrusions is such that the density
structure in front of the intrusion changes in a way which mimics an
enhanced vertical diffusion adjustment. Ivey (1987) has summarized this
work and inferred an ocean averaged diapycnal eddy diffusivity of

FIG.34. The Froude versus Reynolds number plot containing the data from microstruc-
ture segments with well-defined spectral components protruding above a step gradient form.
All dissipations used in this plot were estimated by fitting a Batchelor spectrum to the
spectra computed for the whole segment.
0 Bubble plume thermal PT
0 Thermal falling in a weak ambient stratification
0 Thermal impinging on a sharp interface
A Penetrative convection
V Surface buoyant plume roller region
V Surface buoyant plume wake
x Well mixed surface region
+ Strong stratification, strong shear
Strong stratification, weak unknown shear
+ Strong stratification, no known shear
* Gravitational underflow
A Penetrative convection thermals impinging on an interface
0 Intrusion flows.
0-

3-

-
r 6-
t
0 .

9- -19

26 26-30
TEMPERATURE (“C)
0
n/dz (“Ch)
3010”0
- 10’
& (m2/s3)
10.’ lo4
112

TEMPERATURE (“C)
0
dT/dz (“Urn)
--
3010’0
hw
lo4
Ws3)
10‘ loJ
’I2

h i (1406)
7r

I
I,
-3

w
D
. -
-
E
I
6k
0

- 9

I
’ZLi4 ‘
-
I
- l L
26 28 -30 0 3d10’0’”’;06 ‘;‘Iz
TEMPERATURE (“C) dTidz ( T i m ) EuIx ( m Z / s 3 )

FIG.35. Examples of the development of turbulence in the Harding Reservoir at an


extremely low Lake number L , = 2.2. All profiles extend from 0.2 m above the rocky
bottom to the surface. (a) Temperature profile at 1127 hours. (b) Temperature gradient
profile at 1127 hours. (c) Dissipation. Sticks: Wigner-Ville estimate; bar chart: Batchelor
spectra fitting estimate. (d) As (a) at 1317. (e) As (b) at 1317. (f) As (c) at 1317. (9) As (a)
at 1406. (h) As (b) at 1406. (i) As (c) at 1406.
Physical Limnology 421

3x m2s-l for the interior of the ocean. This result was derived from
the formula
Kl
K , =- (10.7)
BL ’
where K is the diffusion coefficient in the benthic boundary layer, 1 is the
turbulent benthic boundary layer thickness, 8 is the bottom slope, and L
is the basin scale.
The other mechanism is breaking internal waves (Cacchione and
Wunsch, 1974; Murota and Hirata, 1979; Murota et al., 1980; Thorpe,
1987; Garrett and Gilbert, 1988; Ivey and Nokes, 1989; Wallace and
Wilkinson, 1988). Internal waves shoal and break as they propagate over
a sloping bottom, especially if the bottom slope is at the same angle as
the rays of the internal wave field (a critical slope). Ivey and Nokes
(1989) found the turbulence in the benthic boundary layer resulting from
the breaking waves to be sustained by shear leading to Richardson
numbers as low as 0.01. For a critical slope, the steady state boundary
layer thickness resulting from the wave breaking was given by
1 = 5I;, (10.8)
where I; is the internal wave amplitude.
The mixing efficiency within this active boundary layer measured as the
flux-Richardson number (buoyancy flux divided by production) reached a
maximum of 20%. The diffusion coefficient K , can now be estimated
from (10.7) and the estimate of the diffusion coefficient in the benthic
boundary layer (Ivey and Nokes, 1989):
K = 0.09 cot2, ( 10.9)
where I; is the wave amplitude and w the wave frequency. Thus, given
(10.7)-(10.9), a recipe for estimations of the vertical diffusion coefficient
due to boundary mixing is now available. However, application still
requires a knowledge of the internal wave spectrum in a lake as a
function of L N, in order to determine what percentage of the bottom is at
critical slope. This is an urgent priority. Further, given the importance of
the first mode seiche, it is not clear that boundary mixing plays the same
important role in lakes as it appears to do in the ocean.
In summary, a great deal of progress has been made in the area of
hypolimnetic mixing. Recent measurements have conclusively shown the
turbulence level in the interior of the lake increased as LN decreases and
422 Jorg Zmberger and John C . Patterson

also that the effective basin average diffusion coefficient correlates to N- "
where a is approximately 1. The role of the benthic boundary layer and
the turbulence due to shear and internal wave breaking has been
quantified in laboratory experiments and these results are now ready to
be tested in field experiments with the data set from Harding Reservoir.
This should determine the functional relationship between the proportion
of transport in the interior compared with that in the benthic boundary
layer as a function of the Lake number LN .

XI. Modeling

The previous sections have given an overview of the seasonal behavior


of lakes and reservoirs, and discussed in detail the major processes that
determine their density structure. The complex interactions of these
processes which produce the overall thermal characteristics are best
studied with computer simulations. The need for these detailed descrip-
tions has been motivated by lake management requirements, which has,
in the last few decades, led to the development of a variety of modeling
techniques for stratified lakes and reservoirs.
Much of this development has occurred under the assumption of
one-dimensionality, where vertical motions are inhibited and transverse
and longitudinal variations are quickly evened out. Even with this great
simplification, it is difficult to model the interaction of a number of
complex processes to predict a density structure, and a number of models
of varying levels of complexity and success have been produced. There
has also been, to a lesser extent, some development of two- and
three-dimensional stratification models; the increasing complexity and
computational requirements have severely limited this development. The
question of circulation models is a separate issue and is not addressed in
this review.
The early one-dimensional models of stratification concerned the
solution of the one-dimensional heat transport equation. Perhaps the first
of these models was the Dake and Harleman (1969) formulation, in
which molecular diffusivity was assumed to be the vertical transport
mechanism that redistributed the heat input by the surface heat fluxes. In
a preliminary version of later developments, this model generated a
surface mixed layer by redistributing density instabilities produced by
surface cooling. The true diffusivity models followed: Orlob and Selna
Physical Limnology 423

(1970), Huber et al. (1972), Sundaram and Rehm (1973), and Markofsky
and Harleman (1973) all developed models that simulated the vertical
transport of heat in the epilimnion by means of a diffusion-like process,
characterized by an eddy difhsivity E,. The definition of E, was the
major difference in these models. For example, Huber et al. (1972)
assumed a constant value of E, , calibrated from a data sequence. Orlob
and Selna (1970) used a constant value of E, in the epilimnion, with a
value exponentially decaying with depth in the hypolimnion, also
calibrated against data. These models were poor representatives of the
actual processes. Sundaram and Rehm (1973) incorporated some of the
physics into the representation by defining an E, that depended on the
stratification and the wind, through an overall Richardson number.
Similar formulations for E, have been suggested by, for example,
Thomas (1975) and Witten and Thomas (1976). Henderson-Sellers (1976)
has further developed these ideas for deriving E, for thermocline models
based on boundary layer theory and a mixing length turbulence closure
scheme.
Although the early models based on diffusivities were moderately
successful and, being relatively insensitive to the value of E,, were
simple to calibrate, Harleman (1982) showed that this was the result of
the dominance of other processes for the vertical transport of heat, and
the insensitivity may not occur in all cases. Further, in all cases,
calibration was required and transfer of the model from one lake or
reservoir to another required recalibration. Even within one lake,
conditions outside the range of calibration could not be modeled with
confidence. More importantly, the description of the turbulent transport
processes by a single bulk eddy transport coefficient meant that the
effects of the individual processes, such as turbulence generated by the
wind, convective mixing, internal waves, boundary mixing, and basin
scale seiching for example, were integrated into a single parameteriza-
tion. The coefficient was particular to both the lake and the combination
of processes acting at the time, and could have no real physical
interpretation. Because of its integrated, non-physical nature, there was
no independent means of verifying the parameter value or dependence.
Finally, the application of these diffusivities to other lake properties, for
example salinity or dissolved oxygen, was not necessarily correct. Fischer
et al. (1979) pointed out that the distribution of different properties may
arise from different combinations of physical processes, and the same
diffusivity may therefore not be appropriate.
424 Jorg Imberger and John C. Patterson

Recent developments for the incorporation of specific processes, for


example Henderson-Sellers (1984a, 1985), have minimized some of these
difficulties, but if an understanding of the lake behavior in terms of its
short time scale dynamics is required, models based entirely on diffusivity
formulations will be inadequate.
This conclusion led to a number of investigators taking a different
approach to modeling the epilimnion region of the reservoir or lake. The
early developments were based on the Kraus and Turner (1967) model of
thermocline formation in the ocean. As detailed in Section IV, this
provided a deepening rate based on a balance between the turbulent
kinetic energy (TKE) produced at the surface by the stirring of the
surface wind and the energy required at the thermocline to entrain water
from below the thermocline. The first lake models to incorporate a
parameterization of this process were those of Stefan and Ford (1975)
and Harleman and Hurley-Octavio (1977). Tucker and Green (1977)
utilized a version of the Kraus-Turner model for the epilimnion with a
diffusive model with a diffusivity that depended on wind speed and
stratification in the hypolimnion. Tucker and Green also included a
dependence on fetch length of the TKE made available for mixing.
The model development by Harleman and Hurley-Octavio (1977) was
expanded on by Hurley-Octavio et al. (1977) and Bloss and Harleman
(1979). The latter incorporated an effective efficiency of the transfer of
TKE from the surface to the base of the mixed layer, based on a
Richardson number formula. This incorporated the storage effects and
dissipation. This model formed the base for the standard MIT model.
These models all solved for the vertical heat transport on a fixed
discretization of the vertical direction; the lake was effectively split into a
number of horizontal layers of uniform property and fixed thickness.
Vertical transport was then achieved by transport across these boundaries
by mixing from the mixed layer model, diffusion from the eddy diffusion
model in the hypolimnion, or vertical advection as the result of inflow
and outflow. This latter process involved the calculation of vertical
velocities, and consequently, potential problems of numerical diffusion
arose.
A different approach was taken by Imberger et al. (1978) in the first
version of the model DYRESM. Here, the layers were Lagrangian in
character, of variable thickness, and able to move vertically. Thus, inflow
and outflow simply changed the thickness of the layers affected, with
those above moved vertically to accommodate the change in storage. The
TKE resulting from penetrative convection was formally incorporated in
Physical Limnology 425

this version. Later versions of DYRESM also included the effects of


interfacial shear, parameterization of the TKE storage, an eddy
diffusivity for the hypolimnion based on the dissipation of the TKE and
the stratification, and the travel time of the inflows (Spigel and Imberger,
1980; Imberger and Patterson, 1981; Imberger, 1982; Spigel et al., 1986).
The clear advantage of these process-based models was the ability to
characterize the processes independently of the reservoir model. Thus,
although the mixed layer models, for example, contain a number of
parameters, these may be identified as efficiencies of the various
processes and their values derived from field or laboratory experiments.
This independence means that the resulting model may be transferred
from lake to lake without recalibration; a failure of the model to predict
observed data is not necessarily an indication of a faulty parameteriza-
tion, but rather an indication of the existence of processes not para-
meterized. Further, process-based models enable the user to determine
the detailed response of the lake to external forcing.
A third class of one-dimensional model also exists, based on a full
solution of the one-dimensional momentum and energy equations incor-
porating a turbulence closure scheme of a level higher than the eddy
diffusivities discussed above. A number of k--E models of this kind have
been developed (for example, Spalding and Svensson (1977), Svensson
(1978), and Sahlberg (1983)). The computational requirements of this
approach for long term simulations, however, are prohibitive.
The application of one-dimensional models is restricted to those lakes
for which the assumptions hold. Water Resources Engineers, Inc. (1969)
proposed an internal Froude number relating the average throughflow
velocity u from inflows and outflows and the stratification

U
Fr =- (11.1)
(g’H)”’ ’
where g’ was based on the density difference between inflow and surface
water, and H was the total depth. For Fr < l/n, they concluded that the
reservoir was sufficiently stratified for the one-dimensional assumption to
hold. Bloss and Harleman (1979) pointed out that the surface fluxes,
particularly the effects of the surface wind, were not represented in this
criterion and suggested that a limitation on wind speed U, be used in
addition. This was

(11.2)
426 Jorg Imberger and John C. Patterson

where d was the mixed layer depth, L the length of the lake, H the total
depth, and Ap and p o the density jump at the thermocline and the mean
hypolimnion density respectively. If this inequality was not satisfied, it
was assumed that the deviation of the thermocline from the horizontal
was sufficient to prevent the one-dimensional assumption from holding.
The numerical parameter was based on earlier empirical data relating
surface set up to wind speed.
This criterion was similar in character to that established by the
analysis of Spigel and Imberger (1980), based on rigorous dynamical
grounds. Here, as discussed in detail in Section IV, the stratification and
wind forcing were related through the Wedderburn number W
(Thompson and Imberger, 1980) in a series of criteria which categorized
the response of the lake to external wind forcing. These criteria, together
with modified inflow and outflow criteria and a parameter that charac-
terized the importance of the earth’s rotation, were discussed in
Patterson et al. (1984). These may be summarized as follows.
The deviation of the mixed layer was characterized by the value of W
(defined in Section IV). For W > 10, the deviation was negligible, and
the mixed layer deepening was characterized by the one-dimensional
effects of surface stirring. For 3 < W < 10, increasing evidence of shear
production was likely, however the one-dimensional model, with a shear
production algorithm was still valid. For W <3, the interface was
sufficiently close to the surface at the upwind end to be affected by
surface stirring, and the lake became mixed through an upwelling front
moving downwind. These criteria were based on the first mode wave
behavior in a rectangular basin containing a fluid with two densities.
Even with this simplification, the criteria proved a useful classification of
lakes (Patterson et al., 1984).
The inflow criterion was effectively the original Water Resources
Engineers form above (11.1). The approximate nature of the calculation
of Fr meant that there was little significance in the factor l/n, and Fr < 1
was taken to indicate one-dimensional behavior. Similarly, an outflow
Froude number was formed as

F, = Q (11.3)
H2(g’H)”2 ’
where Q was the discharge, which again reflected the relative importance
of inertial and gravitational adjustments. For F, < 1, the one-dimensional
structure was not disturbed. A criterion based on comparing the Rossby
Physical Limnology 427

radius of deformation with the lake scale was used to characterize the
effects of the earth’s rotation.
The difficulty with the Wedderburn number and Froude number
calculations is that they are based on simple geometries and simple
two-layer density structures. To calculate the value of W appropriate to a
real lake involved solving the eigenvalue problem for the internal wave
field in the actual stratification and utilizing this to form an equivalent
two layer structure; depending on the stratification ambiguity could arise.
Likewise, the Froude number criteria were based on very simple
geometries, and took no account of, for example, the ratio of inflow to
storage.
The Lake number LN introduced above minimizes these difficulties of
application. The form discussed in Section I1 has been shown to be
equivalent, in the two layer rectangular case, to W, and its interpretation
is clear. For LN>> 1, the stratification is sufficiently strong to minimize
any disturbance of the metalimnion from the surface wind, and the para-
meterizations of Section IV are appropriate. Thus in an equivalent way to
W, the value of LN characterizes the assumption of the overall one-
dimensionality with respect to surface wind rather than just the surface
layer. Further, by redefining the maximum angle of deviation, the higher
mode responses may also be characterized.
The values of LN calculated for the Canning Reservoir, based on actual
values of u, (rather than the fixed value used in Section 11) and the field
profiles of temperature and salinity, are shown in Figure 36a. This shows
the same trend as Figure 8: LN values of order 50 for the summer period
when the stratification is strong and LN values of order 0-2 in winter
when the stratification is weak and the winds high. Clearly the one-
dimensional assumption is appropriate in summer, but there are periods
during winter when LN is close to 0, and the assumption may be invalid.
Examination of those times, however, reveals that the stratification is
extremely weak, and the error made by a vertical mixing model
compared to a model based on upwelling is small; both produce a
homogeneous reservoir over short time scales. On the other hand, if LN
were small in times of moderate to strong stratification, or W is small but
LN is large, upwelling will occur at the upwind end from the hypolimnion
in the former case and the surface layer in the latter. In both cases, the
one-dimensional assumption does not hold. In these cases, a two- or
three-dimensional circulation model may be more appropriate.
The form of LN for river inflow may also be easily determined.
428 Jorg Imberger and John C. Patterson

+
3w +

1W- ++
+ +
+
+ + t +
50- t
+ + +
+ +++ + + +
+ + + +

3 1.50-
L
9
3 0.00-

ol
0.30

0.00 I I-.
~ *++
I
+
+
+
t

I
+
+
+
t +
I
; k S OI N D J I F M I A I M J I J
MONTH

FIG.36. (a) The Lake number LNfor Canning Reservoir for those days on which field
data are available, based on the actual values of wind speed for the period 11 June 1986 to 7
September 1987. (b) The Lake number LN,[for Canning Reservoir, based on the inflows for
the same period as Figure 36a.
Physical Limnology 429

Following Section X, the Lake number based on inflow can be defined by

(11.4)

where the length scale for the reservoir is taken as Ah”. Thus, large
values of LN,I correspond to small interface deviations.
The values of LN,, for the Canning Reservoir on those days on which
profile data are available are shown in Figure 36b. These are based on
the conservative assumption that zi = 0, that is, that the river underflows.
Clearly, LN.1 increases to relatively large values in the summer when the
stratification, and therefore, S, are large, and the inflows are small. LN,l
becomes small in periods of very weak stratification, and as in the
previous case, the weakness of the one-dimensional assumption causes
little damage.
Using LN and LN,, then gives an indication of the appropriateness of
the one-dimensional assumption for surface and inflow driven deforma-
tions of the density structure. The extension of LN to outflows and
deformations resulting from the earth’s rotations are not obvious and
have not been pursued here.
Given that the indicators for a lake or reservoir suggest that a
one-dimensional parameterization is appropriate and that a process-based
mixing model is preferred to a diffusivity model, the remaining question
is which of the available models should be chosen. As indicated above,
most of the development has been in the context of the MIT and
DYRESM models. In addition, the U.S. Army Corps of Engineers
model CE-QUAL-R1 is available. The temperature prediction com-
ponent of this model is, in its present form (U.S. Army Corps of
Engineers, 1982), very similar to DYRESM. In the following, the
structure and performance of DYRESM only will be discussed in detail.
As noted above, DYRESM is based on a Lagrangian representation of
the lake, with each horizontal layer of uniform property but variable
thickness and location. Thus, each layer expands or contracts as inflow
and outflow affect its volume, and those above move up or down. In this
way all vertical advection of mass is accounted for by layer movement,
and problems of numerical diffusion are not present. Further, conserva-
tion of mass is simpler to achieve without the necessity of computing
vertical velocities.
The model contains five basic process descriptions: surface fluxes of
heat, mass, and momentum; mixed layer dynamics; mixing below the
430 Jorg Imberger and John C . Patterson

surface layer; inflow; and outflow. Each of these is based on the


parameterizations given in the earlier sections and will be described only
briefly. The evolution of the model is described in the literature
(Imberger et al., 1978; Imberger and Patterson, 1981; Spigel and
Imberger, 1980; Imberger, 1982; Patterson et al., 1984; Hocking et al.,
1988). Various extensions to the model are given in Patterson et al.
(1985) (dissolved oxygen), Patterson and Hamblin (1988) (ice cover), and
Ivey and Patterson (1984) (bottom current induced mixing and effect of
rotation on surface layer dynamics).

1. Surface Heat, Mass, and Momentum Fluxes

The fluxes of heat, mass, and momentum at the surface are described
in detail in Section 111. Following the comments in that section with
respect to the self-regulation of the thermal budget, the bulk formulae
equations (3.1)-(3.3) are appropriate for modeling that reports on a daily
time scale, and frequently utilizes meteorological data from a single site,
averaged over a full day. The evaporative heat flux is calculated from
(3.3), using a bulk value of Cw of 2.6 X loF3, within the range of values
suggested. This corresponds to the formulation arising from the Lake
Hefner data (Tennessee Valley Authority, 1972) and is similar to the
form used by Orlob and Selna (1970). Sensible heat transfer (3.2) uses
CH= Cw and is related by the Bowen ratio (Henderson-Sellers, 1986).
The value of CD for momentum transfer (3.1) is wind speed dependent,
in the form (Donelan, 1982; Ivey and Patterson, 1984)
1.124 X lop3, u < 4 m s-', (11.5)
c,=( (0.96 + 0.041 U)x U >4 m s-'.
The remaining heat transfers are treated in the usual way. Incoming
short wave radiation is absorbed by the water column, after reflection at
the surface. The reflection is characterized by the albedo, which is
measured for the particular site. The albedo A, is a function of sun angle,
surface roughness, water color, etc., but is usually taken, for averaged
data, as a constant. Typically, A, is of order 0.04, corresponding to 4%
reflection. The absorption of short wave radiation is characterized by
Beer's Law
(11.6)
Physical Limnology 43 1

where the sum is taken over a number of wavelength bands,

cPi==, (11.7)
i

and the qi are the attenuation coefficients for each band. Although two
and higher band relationships have been used, for example for the
absorption of short wave radiation in ice (Patterson and Hamblin, 1988),
in general a single band formulation is sufficient for use with daily data.
Incoming long wave radiation is either measured or predicted from
formulae such as the Swinbank equation (Swinbank, 1963), modified to
include the effect of cloud (Henderson-Sellers, 1986). Likewise, the long
wave emission from the surface may be measured or calculated from the
Stefan-Boltzmann Law
QL= m T 4 , (11.8)
where E is the surface emissivity, CT the Stefan-Boltzmann constant, and
T the surface temperature in "IS.The emissivity is fixed at 0.975.
These heat transfers are implemented on the layer structure of the
model, with all fluxes except short wave affecting the surface layer only.
The time step of the heating and surface mixing calculations is also set
here; the time is limited to that in which the surface layer temperature
changes by 3"C, up to a maximum of 12 hours. An additional limitation is
placed on the time step in terms of momentum transfer; the change in the
mean mixed layer velocity from the previous value is also limited.

2. Mixed Layer Dynamics

The mixed layer dynamics is modeled by an algorithm that represents a


simplified form of (4.18)-(4.22). Specifically, since the use of daily data
precludes rapid (within a diurnal time scale) changes, which prompted
the Spigel et al. (1986) formulation, the assumption that the turbulence in
the mixed layer adjusts rapidly to changing external inputs is made. This
is equivalent to putting (4.19) equal to zero. Equation (4.18) then
collapses onto the form used in DYRESM,

+--
Apgh
2p0 24p0 dh
+--I
gb2 d ( A p ) g A p b d b dh
12p0 dh dt
CK
=-(w:
2
+ q3u:)+- u:+--+--
"'2[
U f d d U,dd(U,)
6 dh 3 dh 1;'
dh
(11.9)
432 Jorg Imberger and John C. Patterson

where
q: = w: + q3u:, (11.10)
and the following equivalences apply:
77 = CN, (11.11)

(11.12)

(11.13)

CT (CF + cE)-u3. (11.14)


The values used are r,~= 1.23, C, = 0.2, CT= 0.51, and CK= 0.125. These
yield values of CF and CE similar to those given in Section IV. Equation
(4.22) is used to calculate the shear velocity.

3. Mixing below the Surface Layer

As discussed in Section X, mixing below the surface mixed layer is


characterized by patchy, sporadic, individual events that span regions of
only a few meters and last for only a few minutes. These events may arise
from, for example wave-wave interaction, double diffusion, boundary
mixing, or Kelvin-Helmholtz billows. It is not possible, in the context of
a model such as DYRESM, to model individually each of these processes
on the appropriate time and length scales. Rather, the turbulent
processes are modeled by an eddy diffusivity. This approach has been
commonly followed with a wide range of field determinations of K , (see
Section X).
DYRESM utilizes the formulation given by (10.2), as described by
Imberger (1982). Here, the energy inputs from the surface wind and the
inflow are assumed to be uniformly distributed over the surface layer and
the thermocline region. The wave characteristics are determined by
matching the K , with extreme values. Thus, for the case where wind
power input exceeds river inflow, K , from (10.2) is matched with
0.067(H - zT)u*, assuming that u u,. - This yields -
k,
12.4A,/[V‘(H-zo)], where V’ is the volume over which the energy is
dissipated, the lower boundary of which is at zo. Similarly, for the inflow
dominated case, ko-2x/do, where do is the depth of underflow, and
u-0.1 Qldgtan a’,where Q is the flowrate and a’ the stream angle.
Physical Limnology 433

The parameter a in (10.2) is given the value 0.5 (Imberger and Hamblin,
1982).

4. Inflow

The inflow of the rivers into the reservoir is modeled in three separate
processes, as described in Section IX: determination of the plunge point,
underflow, and intrusion. Following Section IX, the plunge point
determination is parameterized by (9.1) and (9.3), the underflow by
(9.2), and the intrusions by (9.4), (9.9, and (9.7)-(9.11). The implemen-
tation of (9.4) and (9.5) is described in Imberger and Patterson (1981);
-
briefly the balance is changed from inertia-buoyancy at R 1 (R defined
by (9.7)). The cases of steep slopes, unconfined, or broad river valleys
are not modeled by DYRESM. The impact of a downflow on a sudden
density change is treated by the method of calculation of N2, as for the
withdrawal case below.
The downflow is modeled by tracking the day’s inflow down the slope
and locating its position and flowing depth. Thus, the underfiow is made
up of a number of parcels of inflows placed along the slope, moving down
on each day until their level of neutral buoyancy is reached. Insertion
following (9.4) or (9.5) generates the intrusion thickness from conserva-
tion of volume. The insertion lengths calculated from (9.4) and (9.5)
together with the downflow characteristics are essentially two-
dimensional parameterizations of the inflow; these are later utilized in a
quasi two-dimensional version (see below).

5. outfiow

In the context of the assumptions of both one- and two-dimensional


forms of DYRESM, the effects of the earth’s rotation are not considered
in the description of the process of withdrawal. Similarly, the behavior in
two-layer fluids is not explicitly modeled, with the behavior relevant to
sudden changes described in Section VIII being modeled in the use of
density gradients averaged across the interface. Although this procedure
is not strictly correct, the error is small. The averaging procedure is
described fully in Imberger and Patterson (1981). The outflow algorithm
is therefore based on (8.23) to (8.26), describing point and line sink
withdrawal from a continuously stratified fluid. Briefly, these equations
are implemented as follows. The inertial withdrawal layer thickness 6
434 Jorg Imberger and John C. Patterson

corresponding to the point source in a stratified fluid is calculated


according to (8.26), with a value of C1of 2, as discussed below. If 6 > W,
where W is the lake width, the layer has spread the width of the lake, and
for all but this immediate spreading region, the layer behaves as though it
is generated by a line sink across the full width of the lake (Lawrence and
Imberger, 1979).
For the point sink case, the inertial limit is described by (8.26), and the
viscous limit by (8.25). The transition flow is modeled, with the
distinction between the two cases being determined by the value of the
parameter S (8.28), following Cases A and B. The species convection
cases described by Ivey and Blake (1985) are not covered by the
algorithm.
If the line sink model is invoked by the condition referred to above,
the layer thickness is described by (8.23) and (8.24), depending on the
value of the parameter RL, as described in Section VIII. The value of C1
for the inertial point sink case is chosen so that the inertial layer thickness
of the point sink case will equal the inertial layer thickness arising from
the line sink, at the discharge at which the transition from point to line
sink occurs. Although slightly larger than the experimental determina-
tions of C1, the error is small and is accepted to ensure a smooth
transition from one type to the other.
The region from which the fluid is drawn is determined by the
algorithm developed by Hocking et al. (1988). Here, a velocity profile is
assumed, following Spigel and Farrant (1984), to be

t uo(1-); [1+ cos( n 3


1, z
0 s -.c1,
6
- 2,
112
(11.15)
z - 2,
0, -2 1,
61/2

which is a cosine profile about the level of the sink with a linear decay
away from the sink to the far end of the lake, where L is the lake length,
dIn the layer half thickness, uo the centerline velocity, and z, the height
of the sink.
Invoking conservation of volume enables computation of the vertical
velocities, and the streamlines were shown by Hocking et al. (1988) to
have the form
Physical Limnology 435

where Y is a constant. Hocking et al. integrated along the streamlines


from the sink to obtain the envelope from which the withdrawal in a
given time step would be taken. Although closed form solutions were not
available, approximate forms for the curves for both large and small
withdrawal were found. The results were verified by comparison with the
Spigel and Farrant (1984) experimental results.
The calculation of these withdrawal envelopes gave an accurate
representation of the quantity of fluid taken from each of the model
layers. This allowed for a proper representation of the horizontal
transport in the withdrawal layer, noting that the envelope extended
above the calculated withdrawal layer near the sink; fluid from outside
the withdrawal layer translated vertically into the withdrawal layer before
moving horizontally into the sink.

These process descriptions are linked together in the framework of


Lagrangian layers introduced above. Thus, mixing in the mixed layer
algorithm takes place by the amalgamation of layers with redistribution
of the layer properties across the amalgamated layer according to
conservation of mass and energy. The thickness of the new layer will
equal the sum of the thicknesses of the individual layers; the lost
resolution is acceptable since there is no variability within the amalgam-
ated layer. Before a process affects that layer, for example heat fluxes at
the surface, the layer is checked against a maximum thickness criterion
and split if necessary. In other words, only the spatial resolution required
for each process is retained. In high gradient regions, the resolution is
correspondingly fine. The temporal resolution is set by the limitations on
heat and mass fluxes described above; again the resolution is fine only in
periods of relatively rapid change. The result is resolution of length scales
down to a few centimeters and time scales down to 15 minutes, but only
when necessary. This means a computationally efficient and accurate
code.
The quality of a simulation by any model depends largely on the
quality of the meteorological, inflow, and withdrawal data. As an
example of the application of DYRESM with relatively good data, a
simulation of the Canning Reservoir, described above, for the period 11
June 1986 to 7 September 1987, (450 days, covering a full seasonal cycle)
has been carried out. The result, presented as isotherm-depth histories, is
shown in Figure 37. This result should be directly compared with Figure
2. Clearly, the simulation has almost exactly followed the data, particu-
436 Jorg Imberger and John C . Patterson
TEMPERATURE ("C

50.0

40.0

- 30.0
N

20.0

10.0

I I I I I I I J I I I I I I
J A S O N D J F M A M J J A

DAY

FIG. 37. Isotherm-depth histories from the simulation of Canning Reservoir by


DYRESM for the period 11 June 1986 to 7 September 1987. This figure should be
compared with Figure 2, which shows the data for the corresponding period.

larly with respect to formation of the stratification through summer and


the subsequent mixed layer deepening in mid-May. There are two points
of mismatch. First, the shape of the 12°C isotherm in the deeper part of
the reservoir differs from the observed data. The error is minor,
however, as the temperature gradients in this region are extremely weak
and small deviations shift the isotherm large distances. Second, the
surface temperature early in the summer period is simulated as being
slightly higher than the data indicate. The error is within 1"C, however.
As a second example of the use of the model for longer term
simulations, an eight-year data set for the Wellington Reservoir was
assembled. Simulations of part of this period have been presented
Physical Limnology 437

d 1914 1975 1978 1 gn 1978 1979 1980 1981 1982


TIME Pmmo

'7

0
1974 1975 1978 wn 1978 1979 1980 1981 1982
T I M PERKID

FIG. 38. The offtake salinities at Wellington Reservoir, measured (dashed line) and
predicted by DYRESM (dotted line), over the extended period 1974 to 1982: (a) mid-level
offtake and (b) scour offtake.

elsewhere (Imberger and Patterson, 1981); here, comparisons of the


predicted offtake properties from the two operating offtakes with the
measured values are shown (Figure 38). Clearly, there are some
deviations from the measured values, but over the long term the
agreement is remarkably good.
These applications show the strength of the one-dimensional version of
DYRESM,both for short term (one year) and longer term (eight years)
simulations. In both cases however, only the one-dimensional variations
in the stratifying species may be modeled, consistent with the assumption
of one-dimensionality. In many cases, however, the parameters of
interest may not influence the density, and even if the density field is
one-dimensional, these parameters may exhibit considerable horizontal
variability. Further, the motions in reservoirs and lakes are in general not
one-dimensional; for example, the wind accelerates the surface waters
and induces a return flow as well as tilting the surface, perhaps leading to
upwelling and internal and surface waves; the density gradients estab-
lished by these and other mechanisms such as unequal heat capture may
drive horizontal motions; inflows to a reservoir are usually of a different
density to the surface water and underflow for some distance before
438 Jorg Zmberger and John C. Patterson

being inserted horizontally. All of these two- and three-dimensional


effects are averaged out in a one-dimensional representation and the
details are lost.
Rodi (1987) reviewed the available models for a broad scope of lake
and reservoir problems. He concluded that in three dimensions the
difficulty of adequately resolving the important time and particularly
length scales limited the quantitative application of these models. In two
dimensions the situation is somewhat better, although models with
stratification are of limited application. These take a vertical surface
along the river valley of the lake, and simulate variations and motions in
the longitudinal and vertical directions. The best known of these models
is the LARM model of Edinger and Buchak (1983) (Gordon, 1981; Kim
et al. 1983) and its more recent version GLVHT (Buchak and Edinger,
1984). This model solves the laterally integrated equations of motion on a
finite difference grid, subject to a number of assumptions. Turbulence
closure was with a mixing length model with a Richardson number
correction. Again, however, resolution remains a major difficulty.
A different approach was taken by Jokela and Patterson (1985) with a
quasi two-dimensional version of DYRESM. The formulation considered
the development of horizontal structure only from inflow intrusions along
the previously defined DYRESM horizontal layers. To identify the
horizontal gradients, each horizontal layer was divided into a number of
Lagrangian parcels that moved horizontally within the layer, changing
length as the layer thickness changed, retaining the Lagrangian character
of the model. The total inflow for the day was distributed across a
number of horizontal layers in the usual way, with each layer’s
apportionment forming a new parcel for that layer. The length of the
parcel was determined from the calculation of the intrusion distance, and
all previous parcels in the layer were forced forward, contracting in
length as the layer thickness increased.
This preliminary model had no provision for adjustment of horizontal
gradients, and was applicable only for situations in which the one-
dimensional assumption held. However, the simulation of an inflow event
showed good agreement with the measured data.
An extended version of this model, with attention paid to other
processes besides inflow, has been completed by Hocking and Patterson
(1988). This model builds on the original Jokela and Patterson model,
including a gravitational adjustment. The vertical mixing processes from
the original one-dimensional model are retained, but are applied on a
parcel by parcel basis.
Physical Limnology 439
a SALINITY (ppt)

--
E
N
30-

20-

1C-

1000 2000 3000 4000 5000 6000 7000

ACCUMULATED DISTANCE (m)

b SALINITY (ppt)

ACCUMULATED DISTANCE (m)

FIG. 39. (a) The isohalines resulting from salinity measurements at a sequence of
stations in Canning Reservoir on 27 October 1986. (b) The isohalines drawn from the
simulation of Canning Reservoir by two-dimensional DYRESM, after a short (30-day)
simulation.
440 Jorg tmberger and John C . Patterson

A simulation of an inflow event in the Canning Reservoir is compared


with the field data in Figure 39, following a 30-day simulation. The
intrusion is characterized by the isohalines shown, salinity being
effectively a tracer in the Canning Reservoir. The simulated intrusion is a
good representation of the field data in terms of the character of the
intrusion and the actual salinity values. The depth and progress of the
intrusion into the reservoir are accurately predicted, and the shape is
qualitatively the same. The thickness is slightly overpredicted. The tilting
of the 0.16ppt isohaline in the deeper part of the reservoir is also well
modeled.
The advantages of this model over those which solve some form of the
equations of motion on a fixed grid are obvious. The horizontal
resolution of the model is determined by the horizontal process acting
and is only fine where required. This means that the algorithm is
computationally economical and the dispersion problems associated with
fixed grids are dispensed with. Further, the tracking of tracers through
the reservoir is made particularly simple by the Lagrangian nature of the
representation.

XII. Reservoir Destratification by Bubble Aerators

The development of modeling techniques for lakes and reservoirs from


detailed process descriptions resulted from an increasing water resource
management requirement. For example, it may be desired to operate a
reservoir to optimize the quality of the withdrawn water, to minimize
evaporation, to prevent long-term build up of some dissolved substance
such as salt, or to prevent algae growth. To design operational strategies
to satisfy such diverse requirements, a process-based model of the kind
described above is essential. Only then, within the limitations of the
assumptions of the model, can the strategies be tested with the processes
properly represented.
As a case in point, reservoir destratification systems are frequently
installed to prevent the build up of stratification, or dismantle an existing
stratification in a reservoir, usually to improve water quality. In summer
the metalimnion is a barrier between the actively mixed epilimnion and
the hypolimnion. The hypolimnion may become deoxygenated, causing
water quality problems. Removing the stratification artificially or directly
aerating the hypolimnion have been practiced for several decades to
Physical Limnology 441

combat this problem. The use of these management techniques in


reservoirs apparently dates from 1954 (Cooley and Harris, 1954),
although applications of bubble curtains as artificial breakwaters existed
in 1907 (see Bulson (1967) and Wilkinson (1979)). A number of reviews
have outlined the means available for destratification (Tolland, 1977;
Henderson-Sellers, 1982), including hydraulic pumping systems, mechan-
ical mixers, and air bubbler systems.
For reservoirs, the most common destratification device is an air
bubbler system. Compressed air is pumped to the bottom of the reservoir
and released; the buoyant plume rises to the surface, carrying with it
water from the hypolimnion, which is then ejected from the plume where
the density of the air-water mixture is approximately equal to the
ambient density. Once ejected, the fluid is locally heavy, and mixing
occurs, reducing the stratification. The release of the air occurs either as
one or more distinct plumes, or as a curtain, depending on the design.
The success of this procedure depends on the bubble plume being
buoyant enough to carry the hypolimnion water into the epilimnion
before discharge. The efficiency therefore depends on both the stratifica-
tion and the air discharge rate.
A typical bubble plume operation was recently documented in the
Harding Reservoir and is shown in Figure 40(a) and (b). In both
examples the wind was negligible, but during the acquisition of the data
shown in Figure 40(a) (see p. 448) the water surface layer was only very
weakly stratified. As evidenced by the shape of the isotherms the bubbler
was sufficiently active (PN = 0.3, see (12.1)) to completely penetrate the
whole water column and yet the stratification was sufficient to return the
cold surface water over the bubbler back to a depth of 6 m ; the falling
plume being nearly 30 m wide.
By contrast the data illustrated in Figure 40(b) was collected during a
windless hot afternoon when the stratification was much stronger at the
surface. Again the isotherms display the impact of the bubbler; the cold
water rising in the plume could not completely penetrate the surface
stratification (PN = 800 for the surface stratification) and the stronger
stratification led to a much thinner plume.
In spite of their relatively long period of use, the design of bubbler
systems falls far short of ideal, with much of the existing design based on
empirical rules determined for a particular water body. For example,
Davis (1980) gives a design procedure based on the stability of the water
body and the energy flux through the surface. Based on a calculation of
442 Jorg Imberger and John C . Patterson

the stability (2.3), the design procedure assumes an efficiency of 5% to


determine the energy input required by the destratifying device. Simi-
larly, determination of the length of diffuser pipe, the diffuser hole
diameters, and the diffuser separation are empirically based. The energy
requirement is based on the strongest stratification and no wind effects,
and is therefore overestimated. The calculations are sensitive to the value
of S,, and the resulting design depends strongly on the initial assumptions
made. Applications using these or similar procedures are described by
Brim and Beard (1980), Brown et al. (1982), and Brady et al. (1983). The
latter demonstrated the sensitivity of the Davis procedure to the assumed
initial profile. Burns (1977) describes some hydraulic model tests in-
tended to assist in the design procedure.
Even though the design procedures are inadequate for lake and
reservoir destratification systems, the bulk of research into bubble plumes
has been done in the context of a homogeneous environment (Kobus,
1968; Wilkinson, 1979; Milgram, 1983; Tacke el al., 1985; Sun and Faeth,
1986; Cheung and Epstein, 1987). These papers describe various aspects
of the plume characteristics, including determinations of the velocity
distribution, entrainment coefficient, and bubble slip velocity. In general
these are based on an integral model of the plume dynamics, with the
exception of the Sun and Faeth (1986) paper, which deals with the
turbulent properties of the plume.
Ditmars and Cederwall (1974), Milgram (1983), and Cheung and
Epstein (1987) presented integral models based on a homogeneous
ambient fluid. The conclusion drawn by these papers has been that, first,
it is not appropriate to treat the plume as a single phase flow unless the
aidow rate is very small, and second, that it is possible to treat the plume
as a simple plume with the buoyancy term modified by the presence of
the bubbles, provided that the slip velocity of the bubbles is
incorporated.
Stratified surrounding fluid was introduced in the models of McDougall
(1978) and Hussain and Narang (1984). The McDougall paper deals with
experiments in a strongly linearly stratified environment and shows that
the behavior of the plume is somewhat more complex than in the
unstratified case. Briefly, the plume rises, carrying relatively heavy water
and entraining ambient fluid, to the point where the mixture density is
approximately equal to the ambient, where the fluid detrains from the
plume. At this point, the detrained fluid has lost the additional buoyancy
of the bubbles and is locally heavy. It therefore plunges to a new level
Physical Lirnnology 443

before spreading horizontally. The bubble plume continues to rise,


creating a new buoyant plume. In a strongly stratified fluid, this process
may occur a number of times, resulting in multiple intrusions, the spacing
of which depends on the stratification and the buoyancy flux. McDougall
proposes two integral models for the plume; the first based on the simple
plume model modified by the buoyancy flux input by the bubbles and the
bubble slip velocity; the second based on a double plume structure, with
an interior plume in which all the bubbles are contained, and surrounded
by an annular outer plume containing only liquid.
Hussain and Narang (1984) developed a similar double plume structure
model for application in weakly stratified fluids. The condition placed on
the stratification implicitly meant that multiple intrusions were not
possible.
In both stratified and unstratified experimental and model studies little
attention had been placed on the behavior under widely varying
stratification conditions and airflow rates. The paper by Asaeda and
Imberger (1988) attempted to resolve this question by conducting
experiments in both two layer and linearly stratified fluids of varying
degrees and varying airflow rates.
Asaeda and Imberger (1988) confirmed experimentally that the
efficiency of the destratification depended on both the stratification and
the airflow rate. For the linear stratification, they showed that the
behavior of a single plume, or, more accurately, the conversion of the
energy input by the plume into mixing, depended on the parameter

(12.1)

where N is the buoyancy frequency, H the depth of the aerator, and Qo


the airflow rate. In particular, they showed that the conversion efficiency
reached a maximum for a value of PN at approximately lo3. The efficiency
was calculated as

(12.2)

representing the ratio of the change in potential energy g AS, in the


stratification to the energy input by the plume in time At (Tolland, 1977).
The maximum efficiency observed was approximately 0.12. The maxi-
mum was also related to the character of the intrusions. For small values
of PN (relative to lo3), the plume travelled to the surface before
444 Jorg Zmberger and John C. Patterson

detraining, similar to the homogeneous experiments of Milgram (1983)


and others; for large values, multiple, unsteady intrusions formed. With
PN near lo3, the plume detrained at one or more locations, with steady
subsurface intrusions forming.
For a two-layer fluid, similar results were obtained. For a given
stratification, low airflow rates produced a plume which detrained at or
below the level of the interface, with the result that very little exchange
between upper and lower layers occurred. At high rates, the detrainment
occurred at the surface and plunged back to the interface. The critical
parameter here was

(12.3)

For PA < 30, the plume broke through the interface and mixed with the
upper layer. For PA > 30, the stratification was strong enough to prevent
the entrained fluid from penetrating the interface, and detrainment
occurred at or below the interface. In the first case, mixing occurred from
above, with a deepening of the interface; in the second, mixing was
weak, and characterized by a thickening of the interface.
These results were consistent with others; the McDougall (1978)
experimental results were of moderate PN, with steady subsurface
intrusions. The two layer results were similar to those obtained by
Graham (1980). While insufficient details are given by Graham to
calculate P A , it is clear that in the early part of the experiments, the
“diffusive” profiles are characteristic of PA > 30; the density jump at the
interface is sufficiently strong to force detrainment below the interface.
At later times, the stratification has been reduced to allow the effective
PA<30, and the plume penetrates fully, resulting in a sharper interface
as mixing from above deepens the upper layer. The experiments of
Kranenburg (1979) may also be put in the context of Pa. The series of
profiles shown in Kranenburg indicate, after the initial stages, a rapidly
deepening interface, consistent with a calculated value of PA of 0.1, much
less than the suggested value of 30 for penetration of the interface. This is
consistent with the qualitative description given by Kranenburg of an
intrusion plunging from the surface to the interface. The experiments of
Dortch and Holland (1980) were directed at comparisons with hydraulic
methods and insufficient details were given for a comparison to be made.
These conclusions have been incorporated into the reservoir model
described above. Although integral models of the plume itself exist, none
Physical Limnology 445

have been coupled to a far field model, other than that of Kranenburg
(1979). Kranenburg developed a three component model; the plume, the
intrusion, and the far field two-layer stratification, in a cylindrical
geometry. Good comparisons with the experimental results were ob-
tained, but the model was not intended for reservoir applications.
The development of the coupled bubble plume and reservoir model is
fully described in Patterson and Imberger (1989); briefly, the single
plume model of McDougall (1978) was discretized using the layer
structure determined by the reservoir model, with the plume entraining
from each layer as it passes. As the plume rises, the effective buoyancy
anomaly decreases as entrainment lowers the plume density and the
ambient density decreases. At some level, the vertical velocity of the
entrained fluid becomes zero, the fluid is ejected and mixes with the
underlying reservoir water. It is assumed that the resulting horizontal
gravitational adjustment is sufficiently rapid for the one-dimensional
adjustment to remain in force. The bubble plume continues, starting a
new entrainment cycle.
The coupled model was verified against a one-year data series from
Myponga Reservoir in South Australia in which an aerator had been
operated over the summer at a depth of 14m below the surface
(Patterson and Imberger, 1989). The model in general reproduced the
resulting thermal structure, and it was concluded that it adequately
represented the coupling of the bubble plume dynamics with the
processes acting in the reservoir.
This enabled the use of the coupled model to determine various aspects
of plume behavior. Of particular interest was the maximum efficiency
airflow rate found in the Asaeda and Imberger (1988) experiments for a
simple geometry and simple stratifications. The question of the ap-
plicability of this to real geometries and stratifications is fundamental to
the optimum design of destratification systems.
One measure of performance of a destratification system is the
efficiency q, defined above. Asaeda and Imberger (1988) showed that the
efficiency depended on the parameters PN and Pa for the two specific
stratifications described. For the more general case the stability of the
system is characterized by S, and the airflow rate by Q,.Similar to the
cases for wind stirring and inflow, it is possible to form a version of the
Lake number appropriate to the bubble plume. Again, following Section
I1 and taking moments about the center of volume yields Stg/3, where /3 is
the angle of deviation of an isopycnal. If the disturbing force is taken as
446 Jorg Imberger and John C. Patterson

the bubble plume, characterized by a friction velocity w , , the moment of


the force scales with pow;H2, since the diameter of the plume scales with
H. The moment is therefore pow:H2L, where L is the length from the
plume to the center of volume. The work done by the plume is
characterized by poQoHg At (Dortch and Holland, 1980); this must be
equivalent to the work done by the disturbing force. Thus,
pow;H2w, At - poQoHg At (12.4)
or
113
w*-(%) . (12.5)

In the same way as in Section X,a generalized number may be defined:

(12.6)

where Bmax is the maximum deviation. For the two-layer case Bmax-
(2, - zT)/L;for the linearly stratified case & , - H / L . In general, if LN,B
is large, the deviation is small compared with Bmax and the density
structure will not be disturbed by the bubble plume; if LN,B is small, the
deviation is large. In the first case, the airflow rate is insufficient to
penetrate the stratification and cause mixing. The second case cor-
responds to an airflow rate well in excess of that required to penetrate the
stratification. If LN,B is large the efficiency will clearly be small since very
little mixing in a relatively strongly stratified environment will occur; if
LN,B is small, although mixing is occurring, the disturbance is more than
is required, and the efficiency is again low. The manifestation of the small
LN,B case is a plume detraining at the surface, which would also have
detrained at the surface with a much lower airflow rate. This qualitative
argument indicates that a maximum value of efficiency should occur at
intermediate values of LN,B and that, following the experiments of
Asaeda and Imberger (1988), the achievable efficiency should be of the
order of 0.12.
The parameters derived by Asaeda and Imberger (1988) are directly
related to L N , B . It is straightforward to show, using the definition of Bmax
above, that, for the linear stratification,

(12.7)
Physical Limnology 447

where A. is the surface area of the experimental tank, and for the two
layer case,

(12.8)

-
For the linear stratification experiments, the value of PN lo3 gives
LN,B -8. For the two-layer experiments that yielded a transition of
character at PA-30, this puts transition at LN,B-5. The definition of
LN,B then unifies the two parameters defined by the experiments of
Asaeda and Imberger, and suggests that for other stratifications, LN,B -
5-8 will maximize efficiency of the bubble plume-induced mixing.
The numerical experiments with reservoir data by Patterson and
Imberger (1989) described above yielded single plume efficiency curves
which indicated a maximum efficiency of order 0.15, consistent with the
Asaeda and Imberger experiments. Patterson and Imberger introduced
the parameter P, where

(12.9)

and M is the total mass of the reservoir, derived from the ratio of energy
stored in the stratification to the energy input by the plume.
The simulations in Patterson and Imberger (1989) showed that a
maximum efficiency occurred at P - 1, as shown in Figure 41a for
Myponga Reservoir. Over the simulation period, a wide range of
stratification is present, and by varying Qo , a wide range of P values may
be obtained. Figure 41 is constructed from the results of several
simulations, with the efficiency plotted as a function of the P calculated
for a particular stratification and airflow without regard to the temporal
order of occurrence.
Myponga Reservoir is a small reservoir of storage 26.8 x lo6m3 and
surface area 2.8 x 106m2.The same calculation has been carried out for
the Harding Reservoir in the northwest of Western Australia. The
Harding Reservoir is substantially larger (storage 63.8 x lo6 m3 and
surface area 14.1 x lo6m2) and is exposed to quite different meteorologi-
cal forces, being located at latitude 20"s (compared with Myponga
Reservoir at 35"s). The calculation of P, however, yields a maximum
-
efficiency again at P 1, with a peak efficiency of 0.15 (Figure 41b). This
parameter therefore unifies the simulation data.
448 Jorg Imberger and John C . Patterson

I TEMPERATURE ("C)
0.0

2.0

--
E
4.0

E
0

6.0

8.0

10.0 I I I
21 I 40.0 60.0 80.0 100.0 120.0

PROJECTED DISTANCE (rn)

FIG.40. Transverse transect across the center of a 400 m long bubble diffuser (1 mm
diameter holes spaced at 4 m) located at a depth of 9 m. Data was obtained by yo-yoing a
temperature probe through the water column at intervals of approximately 8 meters from
the surface to within 1 m of the bottom. Gas flow rate was 40 I s-'. (a) Morning with weak
stratification (April 1989) and (b) Afternoon with strong stratification (April 1989).

The relationship between P and LN,B is given by

For the reservoir, the appropriate Pmax is (H - zT)/L, and the relationship
between L N , B and P is not direct, as both M and zT may change daily.
The calculation of L N , B would require new simulations; these have not
been performed.
The performance of a single plume is characterized by the value of P
and therefore indirectly by ; for optimum performance, a single
-
plume should be operated such that P 1. As the stratification changes,
the airflow rate should change. Further, the degree of destratification
obtained by a plume is given by A(gS,), which will generally be much less
Physical Limnology 449

b TEMPERATURE (OC)

--
E
E
W
n
I

I
28.1
28 1

-280.
./
.BUBBLER
, 280
219-x

PROJECTED DISTANCE (m)

FIG.40. (conrd.)

than gS,. Either more plumes are required or the time over which the
destratification occurs must be extended. An estimate for the number of
-
plumes required is given by n gS,/A(gS,). Patterson and Imberger
(1989), however, show that the efficiency of multiple plumes is higher
than a single plume carrying the same total airflow. The optimum
operational strategy then is to operate as many separate plumes as
-
possible, each with P 1. This of couse involves altering the airflow rate
at perhaps daily intervals, depending on the stratification, a level of
control which may not be possible. In reality, both the number of plumes
and the airflow rate are usually fixed, or variable only in the simplest
sense, such as stopping and starting the aerator.
The effect of this nonvariable operational strategy was demonstrated
by Patterson and Imberger (1989); for Myponga Reservoir, the overall
efficiency achieved was only 0.025 for a 501s-' airflow, the result of
many times when the stratification was weak and the P value extremely
low. This result was strongly dependent on the number of plumes and
450 Jorg Imberger and John C . Patterson
a
20

- 1
I o9 1b-2 1b-1 102

FIG.41. (a) The efficiency of a single bubble plume as a function of the parameter P for
-
Myponga Reservoir, showing a clear maximum at P 1. (b) The efficiency for a single
bubble plume as a function of P for Harding Reservoir. The maximum is again at P 1.-
Physical Limnology 45 1

was achieved with 10 separate plumes. A relatively simple control


strategy, in which the aerator was activated only on those days when the
temperature difference between aerator and surface exceeded 1"C,
doubled the total efficiency to 0.052. This was achieved with 100 separate
plumes. The interaction between the optimum number of plumes, the
total airflow rate, and the operational strategy is a complex one, though it
is clear that further substantial gains in efficiency could be made if control
over the airflow rate is introduced.

XnI. Summary
Advances in physical limnology in the last 10 years, both in the more
traditional area of internal wave seiching and in the newer area of
small-scale motions and mixing, has created a very large discipline. We
have highlighted this progress in this review, and now wish to draw
together those areas which we feel require more work.
(a) Seasonal behavior. Lakes in the tropics may be compared with
those in temperate regions if the stability and forcing are matched. This
may be done by evaluating two nondimensional numbers: the Wedder-
burn number W ,which describes the behavior of the surface layer, and
the Lake number L N , which describes the response of the entire lake.
This hypothesis was formed from a review of earlier work on the
intercomparison of lakes; a detailed intercomparison study of a large
range of lakes in different locations should now be conducted. On the
basis of this it may be found that further nondimensional numbers are
required to specify higher mode responses.
(b) Surface fluxes. A great deal of progress has been made, mainly by
oceanographers, in determining the exchange of momentum, heat, and
water vapor at the air-water interface. Provided the atmospheric internal
boundary layer is well established and the air is either not too stable or
unstable, there are adequate procedures for calculating surface fluxes.
However, few lakes studies have data of sufficient quality to enable the
use of the stability correction. This, together with the well-documented
fact that surface roughness has a strong influence, requires that meteoro-
logical stations be located in the internal boundary layer and that both
stability and roughness be incorporated through a model for the fetch
duration and radiation of waves. Last, the wind is spatially highly
variable over the surface of the lake. This leads to energy input at a
basin-scale length that causes deep hypolimnetic mixing and contributes
452 Jorg Imberger and John C . Patterson

to the strong, horizontal transport processes in the lake, a process so far


completely ignored and one requiring a measure of the wind variability
over the lake.
(c) The su$ace layer. The one-dimensional, vertical energy-based
mixed layer models appear to describe the surface layer well if local wind
and flux data are used. What is now required is a generalization of the
one-dimensional model incorporating the Wedderburn number W , which
governs the behavior of this layer, to account for horizontal advection of
momentum and mass. This should include gravitation-induced advection,
which becomes dominant at the cessation of a wind.
(d) Upwelling. When the Wedderburn number becomes low, the
mixed layer upwells at the upwind end. This upwelling occurs at all values
of the Wedderburn number; it is more severe at low values, but is still
present at values as high as 10 or 20. Recent models of upwelling have
indicated an entrainment process with the same deepening rate as the
rate usually attributed to surface-introduced turbulence. Once again,
these models need to be generalized to allow for two-dimensional effects
and the subsequent horizontal advection spawned at the cessation of
wind.
(e) Differential deepening. This new mechanism describes the uneven
deepening process brought about by the spatial variability of wind stress.
The high degree of variability of the wind over a lake causes extreme
deepening in exposed areas and almost no deepening in sheltered areas,
embayments, and side arms. This observation affects two areas. First, a
lake-averaged surface layer model must somehow account for this
variability, since at the .cessation of each wind period, gravitational
motions even out the isopycnals and the net amount of deepening will
depend on the exact distribution which, in turn, is determined by the
variability of the wind. Second, data from the Wellington Reservoir have
shown that the strong undulations introduced at the base of the mixed
layer by differential deepening lead to the introduction of energy from
the wind through the surface layer directly into the hypolimnion at a
basin scale. This is a much more effective way of getting energy into the
deep part of a lake than with a simple uniform upwelling mechanism
because not only are the first and second modes of the basin excited but
differential deepening energizes higher order internal wave modes. The
relationship between upwelling, basin seiching, and differential deepen-
ing as a function of Wedderburn and Lake numbers must be fully
investigated and the consequences for internal mixing assessed.
Physical Limnology 453
(f) Diflei-enfial heating and cooling. This subject has received a great
deal of attention through the study of the thermal bar in cold lakes but,
because it is the strongest single process for horizontal advection in lakes,
it should receive a great deal more. An assessment of the influence of
frequency of cycling (diurnal) on embayments of various sizes must be
carried out to understand the influence of inertia of the water column.
Only then can we predict the horizontal transport in and out of sidearms.
This is important for both ecology and engineering management, since
recreation in reservoirs expose the lake to pollution. Differential absorp-
tion, the mechanism by which phytoplankton blooms are torn apart by
self-generated convective motions, is a new field. Much work needs to be
done in the interaction of convection and growth dynamics of various
species of plankton.
(g) OutJlow. Selective withdrawal has received a great deal of atten-
tion over the last 20 years and most fundamental flow configurations can
be predicted with reasonable accuracy. The outstanding problems center
around selective withdrawal in a variable stratification, and the influence
of rotation on the initial value problem and the final degeneration into
eddies. Last, the interaction of selective withdrawal and topographic
constraints (sills and contractions) is ready for analysis. This is especially
important since it has been shown that shear waves, which set up the
selective withdrawal layer, are solutions to the Korteweg de Vries
equation, allowing the full interaction to be explored while retaining
some of the nonlinearity of the fluid motion.
(h) lnjlows. An inflow plunges underneath the reservoir water, flows
down the drowned river valley until it reaches a point of neutral
buoyancy, then intrudes into the lake as a horizontal layer. The case
where the entrainment of the underflow is dominated by shear produc-
tion does not seem to have been investigated. Further, the entrainment
properties of the plunge point need urgent attention. The generation of
turbulence or the maintenance of turbulence in the intrusion layer is
important because intrusions can interact with the background internal
wave field to sustain a definite level of turbulent transport. Last,
lake-specific problems such as the interaction of the surface layer with an
inflowing stream and the interaction of the heaving (associated with
internal wave basin scale motions) and the inflow require generalization.
(i) Miring below the surface layer. The central problem here remains.
However, with the introduction of the Lake number, the way to proceed
is clearer. The distribution of turbulent patches and mixing efficiency
454 Jorg Zmberger and John C. Patterson

within each of these patches must be identified as the Wedderburn and


Lake numbers vary. Since, in a typically stratified lake, the volume of
fluid actively involved in turbulent motions at any time is at a maximum
of 1%, this is a most urgent and useful avenue to pursue. Once an
algorithm has been achieved that allows for a statistical description of the
patch distribution and identification of the mechanism underlying each
patch (and thus a description of the mixing efficiency), we will be able to
model deep mixing in a lake. A strong effort should be mounted to
investigate the formation of internal patches and the maintenance of the
turbulent benthic boundary layer.
(j) Modeling. Modeling has proceeded along two avenues: One-
dimensional parametric modeling and full eddy simulation or closure
scheme models. Neither approach is appropriate: two- and three-
dimensional motions are of central importance in a lake, and the scales of
motion range from the Kolmogorov scale to the basin scale. There does,
however, appear to be a spectral gap between the internal waves at the
basin scale and the motions within a turbulent patch. Further, as
mentioned above, the volume occupied by turbulent motions in a lake at
any one time is likely to be less than 1%. A two-dimensional model is
needed that accounts for the mean motions generally described by basin
seiching, superimposed on which is a model that determines the statistical
distribution of turbulence within the lake, in the benthic boundary layer
and surface layer. Last, a further model must be superimposed on these
two that adjusts the density field according to certain mixing efficiencies.
Such a closure scheme has a real chance of success in even quite large
lakes, with present computing power. Many variations are of course
possible, ranging from a simple parameterization of the patch distribution
and the patch efficiency to full turbulent simulations within each patch
once they have been identified.
(k) Reservoir destratijication by bubble aeration. It is unwise to try to
destratify a lake completely by introducing mechanical energy. It is much
more energy efficient to weaken the structure through the bubble flow
and allow the wind to complete the mixing. This requires an examination
of the bubble flow and the natural reservoir dynamics. This area of
research seems to be in its infancy, as does the equivalent research where
propellors are used rather than bubble plumes. Given that the field of
reservoir destratification is an enormous industry, large cost savings are
no doubt possible.
Physical Limnology 455

Acknowledgements
Sections I to X were written by Jorg Imberger and Sections XI and XI1 were written by
John Patterson. We would like to thank Greg Ivey and Graeme Hocking for their
comments on the manuscript. Carolyn Oldham prepared the figures and performed the
computations, and Colleen Henry-Hall edited the manuscript and prepared it for
publication. This assistance is gratefully acknowledged. This work was supported by the
Centre for Environmental Fluid Dynamics, the Australian Research Council, the Australian
Water Research Advisory Council, and the West Australian Water Authority.

References
Akiyama, J., and Stefan, H. G. (1984). Plunging flow into a reservoir: Theory, J . Hydr.
Eng. 110,484-499.
Akiyama, J., and Stefan, H. G. (1985). Turbidity current with erosion and deposition. J .
Hydr. Eng. 111, 1473-1496.
Akiyama, J., and Stefan, H. G. (1987). Gravity currents in lakes, reservoirs and coastal
regions: Two-layer stratified flow analysis. Proj. Rept. No. 253, University of
Minnesota.
Anderson, R. J., and Smith, S. D. (1981). Evaporation coefficient for the sea surface from
eddy flux measurement. J . Geophys. Res. 86, 449-456.
Andre, J. E., and Lecarrere, P. (1985). Mean and turbulent structures of the oceanic
surface layer as determined from one-dimensional third-order simulations. J . Phys.
Oceanogr. 15, 121-132.
Arai, J. (1964). Some relations between the thermal property of lake and its fetch size.
Geogr. Rev. of Japan 37, 131-137.
Arita, M., Jirka, G. H., and Tamai, N. (1986). Classification and mixing of two-
dimensional buoyant surface discharges. 1. Hydr. Res. 24, 333-345.
Asaeda, T., and Imberger, J. (1988). Structures of bubble plumes in stratified environ-
ments. Environmental Dynamics Rept. ED-88-250, Centre for Water Research,
University of Western Australia.
Atkinson, J. F. (1988). Interfacial mixing in stratified flows. J. Hydr. Res. 26, 27-31.
Atkinson, J. F.,and Harleman, D. R. F. (1983). A wind-mixer layer model for solar ponds.
Solar Energy 31,243-259.
Bean, B. R., Emmanuel, C. B., Gilmer, R. O., and McGavin, R. E. (1975). The spatial
and temporal variations of the turbulent fluxes of heat, momentum and water vapor
over Lake Ontario. J. Phys. Oceanogr. 5 , 532-540.
Bejan, A., Al-Homoud, A. A., and Imberger, J. (1981). Experimental study of
high-Rayleigh number convection in a horizontal cavity with different end tempera-
tures. J. Fluid Mech. 109, 283-299.
Bejan, A., and Rossie, A. N. (1981). Natural convection in a horizontal duct connecting
two fluid reservoirs. J . Heat Transfer 103, 108-113.
Bejan, A., and Tien, C. L. (1978). Laminar natural convection heat transfer in a horizontal
cavity with different end temperatures. J . Heat Transfer 100,641-647.
Bella, D. A. (1970). Simulating the effect of sinking and vertical mixing on algal population
dynamics. J. Water Pollution Control Fed. 42(5), 140-152.
Bennett, J. R. (1971). Thermally driven lake currents during the spring and fall transition
periods. In “Proc. 14th C o d . Great Lakes Res.,” Ann Arbor p. 535. Intl. Assoc. for
Great Lakes Res.
456 Jorg Zmberger and John C . Patterson
Benney, D. J. (1966). Long nonlinear waves in fluid flows. J. Math. Phys. Sci. 45, 52-63.
Blake, S., and hey, G. N. (1986a). Density and viscosity gradients in zoned magma
chambers, and their influence on withdrawal dynamics. J . Volcanology and Geothermal
Res. 30,201-230.
Blake, S . , and hey, G. N. (1986). Magma-mixing and the dynamics of withdrawal from
stratified reservoirs. J . Volcanology and Geothermal Res. 27, 153-178.
Blanc, T. V. (1983). A practical approach to flux measurements of long duration in the
marine atmospheric surface layer. J. Clim. Appl. Met. 22, 1093-1110.
Blanc, T. V. (1985). Variation of bulk-derived surface flux, stability, and roughness results
due to the use of different transfer coefficient schemes. J . Phys. Oceanogr. 15,
650-669.
Blanc, T. V. (1987). Accuracy of bulk-method-determined flux, stability and sea surface
roughness. J . Geophys. Res. 92,3867-3876.
Blanton, J. 0.(1973). Vertical entrainment into the epilimnia of stratified lakes. Limnol.
Oceanogr. 18,697-704.
Bloss, S., and Harleman, D. R. F. (1979). Effect of wind-induced mixing on the seasonal
thermocline in lakes and reservoirs. Rept. No. 249, Massachusetts Institute of
Technology.
Bohan, J. P., and Grace, J. L. (1973). Selective withdrawal from man-made lakes. Tech.
Rept. No. H-73-4,Waterways Experiment Station, Vicksburg, Mississippi.
Boyce, F. M., Robertson, D. G., and hey, G. N. (1983). Summer thermal structure of
Lake Ontario off Toronto: Cooling the big city. Atmos. Ocean. 21,397-417.
Brady, J. A., Davis, J. M., and Douglas, E. W.(1983). First filling of Kielder Reservoir
and future operational considerations. J. Inst. Water Eng. Sci. 37, 295-312.
Brim, W. D., and Beard, J. D. (1980). Reservoir management by pneumatic induced
circulation. Paper No. 5-16.In “Proc. Symp. on Surface Water Impoundments,” (H.
G. Stefan, ed.), ASCE, pp. 877-885. Minneapolis, Minnesota.
Britter, R. E., and Simpson, J. E. (1978). Experiments on the dynamics of a gravity current
head. J. Fluid Mech. 88, 223-240.
Britter, R. E., and Simpson, J. E. (1981). A note on the structure of the head of an
intrusive gravity current. J . Fluid Mech. 112,459-466.
Brocard, D. N., and Harleman, D. R. F. (1980). Two-layer model for shallow horizontal
convective circulation. I. Fluid Mech. 100, 129-146.
Browand, F. K.,Guyomar, D., and Yoon, S. C.(1987). The behavior of a turbulent front
in a stratified fluid: Experiments with an oscillating grid. J . Geophys. Res. 92,
5329-5341.
Brown, I. K., Wooley, D. A., and Jory, A. G. (1982). Artificial destratification of Lake
Moms to improve water quality. In “Proc. Hydr. and Water Resources Symp.,”
Institution of Engineers, Australia, Melbourne, Victoria.
Brubaker, J. M. (1987). Similarity structure in the convective boundary layer of a lake.
Nature 330,742-745.
Bryant, P. J., and Wood, I. R. (1976). Selective withdrawal from a layered fluid. J. Fluid
Mech. 77,581-591.
Buchak, E. M., and Edinger, J. E. (1984). Generalized, longitudinal-vertical hydrodynam-
ics and transport: Development, programming and applications. Doc. No. 84-WR,
Waterways Experiment Station, Vicksburg, Mississippi.
Bulson, P. S. (1967).Theory and design of bubble breakwaters. Dock and Harbour Aurh.
48(560), 47-54.
Bunn, S . E., and Edward, D. H. D. (1984). Seasonal meromixis in three hypersaline lakes
on Rottnest Island, Western Australia. Aust. J . Mar. Freshw. Res. 35,261-265.
Physical Limnology 457
Bums, F. L. (1977). Localized destratification of large reservoirs to control discharge
temperatures. Prog. Waf. Tech. 9, 53-63.
Busch, N. E. (1977). Fluxes in the surface boundary layer over the sea. In “Modeling and
Prediction of the Upper Layers of the Ocean” (E. B. Kraus ed.), pp. 72-91, Pergamon
Press, New York.
Businger, J. A. (1955). On the structure of the atmospheric surface layer. J . Meteor. U ,
553-561.
Businger, J. A. (1973). Turbulent transfer in the atmospheric surface layer. In “Workshop
on Micrometeorology” (D. A. Hauger ed.), pp. 67-100. Science Press, Ephrata,
Pennsylvania.
Cacchione, D., and Wunsch, C. (1974). Experimental study of internal waves over a slope.
J, Fluid Mech. 66, 223-239.
Caldwell, D. R. (1983). Oceanic turbulence: Big bangs or continuous creation? 1. Geophys.
Res. 88,7543-7550.
Carmack, E. C. (1986). Circulation and mixing in ice-covered waters. In “The Geophysics
of Sea Ice” (N. Untersteiner ed.), pp. 641-712. Plenum Press, New York.
Carslaw, H. S., and Jaeger, J. C. (1978). Conduction of Heat in Solids, Third Edition.
Oxford University Press.
Carson, D. J., and Richards, P. J. R. (1978). Modeling surface turbulent fluxes in stable
conditions. Boundary Layer Met. 14, 67-81.
Cavaleri, L., and Zecchetto, S. (1987). Reynolds stresses under wind waves. J. Geophys.
Res. 92,3894-3904.
Chao, S . Y. (1988). River-forced estuarine plumes. J. Phys. Oceanogr. 18, 72-88.
Charnock, H. (1955). Wind stress on a water surface. Quart. J . Roy. Meteor. SOC. 82,
639-640.
Chen, J. C. (1980). Studies on gravitational spreading currents, Rept. No. KH-R-40,
California Institute of Technology.
Cheung, F. B., and Epstein, M. (1987). Two-phase gas bubble-liquid boundary layer flow
along vertical and inclined surfaces. Nuclear Eng. and Design. 99, 93-100.
Christodoulou, G. G. (1986). Interfacial mixing in stratified flows. J . Hydr. Res. 24, 77-92.
Chu, V. H., and Baddour, R. E. (1984). Turbulent gravity-stratified shear flows. 1. Fluid
Mech. Us,353-378.
Chu, V. H., and Jirka, G. H. (1985). Surface buoyant jets and plumes. In “Encyclopedia of
Fluid Mechanics” (N. Cheremisinoff ed.), pp. 1053-1084. Gulf Publishing, Houston,
Texas.
Clarke, R. H., Dyer, A. J., Brook, R. R., Reid, D. G., and Troup, A. J. (1971). The
Wangara experiment: Boundary layer data. Tech. Paper No. 19, CSIRO Div. Met.
Phys., Melbourne, Victoria.
Claussen, M. (1987). The flow in a turbulent boundary layer upstream of a change in
surface roughness. Boundary Layer Met. 40,31-86.
Cohen, A. (1979). Existence and regularity for solutions of the Korteweg-de Vries
equation. Arch. Rat. Mech. Anal. 71, 143-175.
Callings, I. L. (1985). Two large Froude numbers cusped free-surface flows due to a
submerged line source or sink. Manuscript. Deakin University, Victoria.
Colman, J. A,, and Armstrong, D. E. (1987). Radon-222 determined vertical eddy
diffusivity in the benthic boundary layer of ice-covered lakes. Limnol. Oceanogr. 32,
577-591.
Cooley, P., and Harris, S. L. (1954). The prevention of stratification in reservoirs. J. Inst.
Water Eng. 8, 517-537.
Cormack, D. E., Leal, L. G., and Imberger, J. (1974). Natural convection in a shallow
458 Jorg Imberger and John C. Patterson
cavity with differentially heated end walls. Part 1: Asymptotic theory. J. Fluid Mech.
65,209-230.
Cormack, D. E., Stone, G. P., and Leal, L. G. (1975). The effect of upper surface
conditions on convection in a shallow cavity with differentially heated end walls. Zntl. J .
Heat Mass Transfer 18,635-648.
Craya, A. (1949). Recherches theoriques sur I’ecoulement de couches superposees de
fluides de densites differentes. La Houille Blanche Jan.-Feb., 44-55.
Csanady, G. T. (1972). Response of large stratified lakes to wind. J . Phys. Oceanogr. 2,
3-13.
Csanady, G . T. (1982). “Circulation in the Coastal Ocean.” D. Reidel Publishing, Boston.
Dake, J. M. K., and Harleman, D. R. F. (1969). Thermal stratification in lakes: Analytical
and laboratory studies. J. Water Resources Res. 5 , 484-495.
Darbyshire, J., and Colclough, M. (1972). Measurement of thermal conductivities in a
freshwater lake. Geofiica Pura i Applicata 93,151-158.
Darden, R. B., Imberger, J., and Fischer, H. (1975). Jet discharge into a stratified
reservoir. J. Hydr. Diu. ASCE 101, 1211-1220.
Davis, J. M. (1980). Destratification of reservoirs-A design approach for perforated-pipe
compressed-air systems. Water Services 84, 497-504.
de Szoeke, R. A. (1980). On the effects of horizontal variability of wind stress on the
dynamics of the ocean mixed layer. J . Phys. Oceanogr. 10,1439-1454.
de Szoeke, R. A., and Rhines, P. B. (1976). Asymptotic regimes in mixed-layer deepening.
J . Marine Res. 34, 111-116.
Deardorff, J. W. (1968). Dependence of air-sea transfer coefficients on bulk stability. J .
Geophys. Res. 73,2549-2557.
Deardorff, J. W. (1970). Convective velocity and temperature scales for the unstable
planetary boundary layer. J. Atinos. Sci. 27, 1211-1213.
Deardorff, J. W. (1974). Three-dimensional numerical study of the height and mean
structure of a heated planetary boundary layer. Boundary Layer Met. 7,81-106.
Debler, W. R. (1959). Stratified flow into a line sink. J . Eng. Mech. Diu. 85, 51-65.
Denton, R. A,, and Wood, I. R. (1981). Penetrative convection at low Peclet number. J .
Fluid Mech. 1l3, 1-21.
Dickey, T. D., and Simpson, J. J. (1983). The influence of optical water type on the diurnal
response of the upper ocean. Tellus 35B, 142-154.
Dillon, T. M., Richman, J. G., Hansen, C. G., and Pearson, M. D. (1981). Near surface
turbulence measurements in a lake. Nature 290, 390-392.
Ditmars, J. D., and Cedenvall, K. (1974). Analysis of air-bubble plumes. In “Proc. Coastal
Eng. Conf.,” Copenhagen. Ch. 128, pp. 2209-2226.
Djordjevic, V. D., and Redekopp, L. G. (1978). The fission and disintegration
- of internal
solitary waves moving over two-dimensional topography. J. Phys. Oceanogr. 8 ,
1016-1024.
Dobson, F. W., Hasse, L., and Davis, R. (1980). “Air-Sea Interaction Instruments and
Methods.” Plenum Press, New York.
Donelan, M. A. (1980). Similarity theory applied to the forecasting of wave heights,
periods and directions. In “Proc. Canadian Coastal Conf. ,” pp. 47-61. Burlington,
Ontario.
Donelan, M. A. (1982). The dependence of the aerodynamic drag coefficient on wave
parameters. In “First Intl. Conf. on Meteorology and Air/Sea Interaction of the
Coastal Zone” (preprint volume), pp. 381-387. The Hague, American Met. SOC.,
Boston.
Donelan, M. A., and Peirson, W. J. (1987). Radar scattering and equilibrium ranges in
Physical Limnology 459

wind-generated waves with application to scatterometry. J . Geophys. Res. 92,


4971-5029.
Dortch, M. S., and Holland, J. P. (1980). Methods of total lake destratification. Paper No.
5-19. In “Proc. Symp. on Surface Water Impoundments,” ASCE, pp. 913-922.
Minneapolis, Minnesota.
Drazin, P. G., and Reid, W. H. (1981). “Hydrodynamic stability.” Cambridge University
Press, New York.
Drummond, J. E., and Korpela, S. A. (1987). Natural convection in a shallow cavity. 1.
Fluid Mech. 182, 543-564.
Dubreil-Jacotin, M. L. (1934). Sur les theorems dexistence relatifs a w ondes permanentes
periodiques a deux dimensions dan les liquids heterogenes. J. de Mathematiques 3,
217-291.
Dyer, A. J. (1974). A review of flux-profile relationships. Boundary Layer Met. 7 , 363-372.
E, X., and Hopfinger, E. J. (1986). On mixing across an interface in stably stratified fluid.
J. Fluid Mech. 166, 227-244.
Edinger, J. E., and Buchak, E. M. (1983). Developments in L A W : A longitudinal-
vertical, time-varying hydrodynamic reservoir model. Tech. Rept. E-83-1, Waterways
Experiment Station, Vicksburg, Mississippi.
Elder, J. W. (1959). The dispersion of marked fluid in turbulent shear flow. J . Fluid Mech.
5,544-560.
Elliott, G. H., and Elliott, J. A. (1970). Laboratory studies on the thermal bar. In “Proc.
13th Conf. Great Lakes Res.,” pp. 413-418. Intl. Assoc. Great Lakes Res. Ann Arbor.
Ellison, T. H., and Turner, J. S. (1959). Mixing of a dense fluid in a turbulent pipe flow.
Part 1: Overall description of the flow. J . Fluid Mech. 33, 514-528.
Emmanuel, C. B. (1975). Drag and bulk aerodynamic coefficients over shallow water.
Boundary Layer Met. 8, 465-474.
Farmer, D. M. (1978). Observations of long nonlinear waves in a lake. J. Phys. Oceanogr.
8, 63-73.
Farmer, D. M., and Cartnack, E. (1982). Wind mixing and restratification in a lake near the
temperature of maximum density. J . Phys. Oceanogr. 11, 1516-1533.
Farrell, G. J., and Stefan, H. G. (1986). Buoyancy induced plunging flow into reservoirs
and coastal regions. Proj. Rept. No. 241, St. Anthony Falls Hydraulic Lab., University
of Minnesota.
Faust, K. M. (1981). Modelldarstellung von Warmeinselstromungen durch Konvek-
tionsstrahlen. Paper No. SFB 80/ET/201, Universitat Karlsruhe.
Findikakis, A. N., and Street, R. L. (1982a). Finite element simulation of stratified
turbulent flows. J. Hydr. Div. ASCE 108, 904-920.
Findikakis, A. N., and Street, R. L. (1982b). Mathematical description of turbulent flows.
I. Hydr. Div. A X E 108, 887-903.
Findikakis, A. N., and Street, R. L. (1983). Approximate numerical model for stratified
flow. 1. Eng. Mech. 109, 950-969.
Fischer, H. B., and Smith, R. D. (1983). Observations of transport to surface waters from a
plunging inflow to Lake Mead. Limnol. Oceanogr. 28, 258-272.
Fischer, H. B., List, E. J., Koh, R. C. Y., Imberger, J., and Brooks, N. H. (1979).
“Mixing in Inland and Coastal Waters.” Academic Press, New York.
Forbes, L. K., and Hocking, G. C. (1988). Flow caused by a point sink in a fluid having a
free surface. J . Aust. Math. SOC. Ser. B (in press).
Ford, D. (1978). Unpublished data. Environ. Lab., Waterways Experiment Station,
Vicksburg, Mississippi.
Ford, D. E., and Johnson, M. C. (1981). Field observations of density currents in
460 Jorg Imberger and John C. Patterson
impoundments. Surface Water Impoundments, H. G. Stefan, ed. ASCE. Minneapolis,
Minnesota. pp. 1239-1248.
Ford, D. E., Johnson, M. C., and Monismith, S. G. (1980). Density inflows to Degray
Lake, Arkansas. In “Second Intl. Symp. on Stratified Flows,” pp. 977-987.
Trondheim.
Franke, R., Leschziner, M. A., and Rodi, W. (1987). Numerical simulation of wind-driven
turbulent flow in stratified water bodies. In “Proc. 3rd Intl. Symp. on Stratified Flows,”
(E. J. List, ed.) ASCE Pasadena, California.
Friehe, C. A., and Schmitt, K. F. (1976). Parameterization of air-sea interface fluxes of
sensible heat and moisture by the bulk aerodynamic formulas. J . Phys. Oceanogr. 6,
801-809.
Fritts, D. C. (1984). Gravity wave saturation in the middle atmosphere: A review of theory
and observations. Rev. Geophys. Space Phys. 22, 275-308.
Gargett, A. E., and Holloway, G. (1984). Dissipation and diffusion by internal waves
breaking. J . Marine Res. 42, 15-27.
Gariel, P. (1949). Recherches experimentales sur I’ecoulement de couches superposees
de fluides de densites differentes. La Houille Blanche Jan.-Feb., 44-55.
Garratt, J. R. (1977). Review of drag coefficients over oceans and continents. Monthly
Weather Review 105,915-929.
Garratt, J. R. (1987). The stably stratified internal boundary layer for steady and diurnally
varying offshore flow. Boundary Layer Met. 38,369-394.
Garratt, J. R., and Ryan, B. F. (1988). The structure of the stably stratified internal
boundary layer in offshore flow over the sea. Boundary Layer Met. 47, 17-40.
Garrett, C., and Gilbert, D. (1988). Estimates of vertical mixing by internal waves reflected
off a sloping bottom. In “Small-Scale Turbulence and Mixing in the Ocean: Proc. 19th
Intl. Leige Coll. on Ocean Hydrodyn.” (J. J. Nihoul and B. M. Jamart eds.), pp.
405-424. Elsevier, New York.
Garrett, C., and Munk, W. (1979). Internal waves in the ocean. Ann. Rev. Fluid Mech. 11,
339-369.
Garwood, R. W. (1977). An oceanic mixed layer model capable of simulating cyclic states.
J . Phys. Oceanogr. 7,455-468.
Geernaert, G. L., and Katsaros, K. B. (1986). Incorporation of stratification effects on the
oceanic roughness length in the derivation of the neutral drag coefficient. J. Phys.
Oceanogr. 16, 1580-1584.
Geernaert, G. L., Katsaros, K. B., and Richter, K. (1986). Variation of the drag coefficient
and its dependence on sea state. J. Geophys. Res. 91,7667-7679.
Geernaert, G. L., Larsen, S. E., and Hansen, F. (1987). Measurements of the wind stress,
heat flux and turbulence intensity during storm conditions over the North Sea. J .
Geophys. Res. 92, 13127-13139.
Gelhar, L. W., and Mascolo, D. M. (1966). Non-diffusive characteristics of slow viscous
stratified flow towards a line sink. Massachusetts Institute of Technology Hydrodynam-
ics Lab. Rept. No. 88, pp. 1-39.
Gibson, C. H. (1980). Fossil temperature, salinity and vorticity turbulence in the ocean. In
“Marine Turbulence” (J. C. J. Nihoul ed.), pp. 221-257. Elsevier, New York.
Gibson, C. H. (1982). On the scaling of vertical temperature gradient spectra. J . Geophys.
Res. 87, 8031-8038.
Gibson, C. H. (1986). Internal waves, fossil turbulence and composite ocean microstructure
spectra. J . Fluid Mech. 168,89-117.
Gibson, C. H. (1987a). Fossil turbulence and intermittency in sampling oceanic mixing
processes. J . Geophys. Res. 92, 5383-5404.
Physical Limnology 46 1
Gibson, C. H. (1987b). Oceanic turbulence: Big bangs and continuous creation. J ,
Physiochem. Hydrodyn. 8, 1-22.
Gibson, C. H. (1988). Evidence and consequences of fossil turbulence in the ocean. In
“Small-Scale Turbulence and Mixing in the Ocean, Proc. 19th Intl. Leige Coll. on
Ocean Hydrodyn.” (J. J. Nihoul and B. M. Jamart, eds.), pp. 319-334. Elsevier, New
York.
Gill, A. E. (1966). The boundary-layer regime for convection in a rectangular cavity. J.
Fluid Mech. 26, 515-536.
Gill, A. E., and Clarke, A. J. (1974). Wind induced upwelling, coastal currents and sea
level changes. Deep-sea Res. 21, 325-345.
Gordon, J. A. (1981). LARM two-dimensional model: An evaluation. J. Enuiron. Eng.
Diu. 107, 877-886.
Graf, W. H., Merzi, N., and Pemnjaquet, C. (1984). Aerodynamic drag measured at a
nearshore platform on Lake Geneva. Arch. Met. Geophys. Bioklim. A33, 151-173.
Graham, D. S. (1980). Destratification characteristics and energy efficiency of air-plume
mixing of stratified fluids. Paper No. 5-20. In “Proc. Symp. on Surface Water
Impoundments,” ASCE pp. 923-932. Minneapolis, Minnesota.
Gregg, M. C. (1987). Diapycnal mixing in the thermocline: A review. J. Geophys. Res. 92,
5249-5286.
Gregg, M. C., and Briscoe, M. G. (1979). Internal waves, finestructure, microstructure and
mixing in the ocean. Rev. Geophys. Space Phys. 17, 1524-1548.
Gregg, M. C., and Sanford, T. B. (1988). The dependence of turbulent dissipation on
stratification in a diffusively stable thermocline. J . Geophys. Res. 93, 12381-12392.
Hannoun, I. A., and List, E. J. (1988). Turbulent mixing at a shear-free density interface.
1. Fluid Mech. 189, 211-234.
Hannoun, I. A., Fernando, H. J. S., and List, E. J. (1988). Turbulence structure near a
sharp density interface. J . Fluid Mech. 189, 189-209.
Harashima, A., and Watanabe, M. (1986). Laboratory experiments on the steady
gravitational circulation excited by cooling of the water surface. J. Geophys. Res. 91,
13056- 13064.
Harleman, D. R. F. (1982). Hydrothermal analysis of lakes and reservoirs. J . Hydr. D i n
ASCE 108,302-325.
Harleman, D. R. F., and Hurley-Octavio, K. A. (1977). Heat transport mechanisms in
lakes and reservoirs. In “Proc. 17th Cong. Intl. Assoc. Hydr. Res.,” Baden
Baden.
Harleman, D. R. F., Morgan, R. L., and Purple, R. A. (1959). Selective withdrawal from a
vertically stratified fluid. In “Proc. 8th Congr. Intl. Assoc. for Hydr. Res.,” Montreal,
pp . 10-C-1-10-C-16.
Hart, J. E. (1972). Stability of thin non-rotating Hadley circulations. J. Amos. Sci. 29,
687-697.
Hart, J. E. (1983a). Low Prandtl number convection between differentially heated end
walls. Intl. J. Heat Mass Transfer 26, 1069-1074.
Hart, J. E. (1983b). A note on the stability of low-Prandl number Hadley circulations. J.
Fluid Mech. l32,271-281.
Hauenstein, W., and Drams, T. H. (1984). Investigation of plunging density currents
generated by inflows in lakes. J . Hydr. Res. 22, 157-179.
Heaps, N. S. (1984). Vertical structure of current in homogeneous and stratified waters. In
“Hydrodynamics of Lakes” (K. Hutter ed.), pp. 153-207. Springer-Verlag. New York.
Heaps, N. S., and Ramsbottom, A. E. (1966). Wind effects on water in a narrow
two-layered lake. Phil. Trans. Roy. SOC.London Ser. A 259, 391-430.
462 Jorg Imberger and John C . Patterson
Heathershaw, A. D., and Martin, J. (1987). The theory and tuning of the Met 0.20 mixed
layer model. Admirality Research Establishment Accn. No. 75636.
Hebbert, R., Imberger, J., h h , I., and Patterson, J. (1979). Collie River underflow into
the Wellington Reservoir. J. Hydr. Div. ASCE 105,533-545.
Henderson-Sellers, B. (1976). Role of eddy diffusivity in thermocline formation. Proc.
ASCE 102,517-531.
Henderson-Sellers, B. (1982). Reservoir destratification techniques and models. In ASCE
“Proc. Applying Res. to Hydraulic Practice Conf. ,” pp. 697-706. Mississippi.
Henderson-Sellers, B. (1984a). Development and application of “U.S.E.D.”: A hydroclim-
ate lake stratification model. Ecol. Modelling 21,233-246.
Henderson-Sellers, B. (1984b). “Engineering Limnology.” Pitman, Marshfield,
Massachusetts.
Henderson-Sellers, B. (1985). New formulation of eddy diffusion thermocline models.
Appl, Math. Modelling 9, 441-446.
Henderson-Sellers, B. (1986). Calculating the surface energy balance for lake and reservoir
modeling: A review. Rev. Geophys. 24, 625-649.
Hicks, B. B. (1972). Some evaluations of drag and bulk transfer coefficients over water
bodies of different sizes. Boundary Layer Met. 3, 201-213.
Hicks, B. B. (1975). A procedure for the formulation of bulk transfer coefficients over
water. Boundary Layer Met. 8, 315-324.
Hicks, B. B. (1975). Wind profile relationships from the ‘Wangara’ experiment. Quart. J.
Roy. Met. SOC.102, 535-551.
Hicks, B. B., Drinklow, R. L., and Grauze, G. (1974). Drag and bulk transfer coefficients
associated with a shallow water surface. Boundary Layer Met. 6,287-297.
Hocking, G. C. (1985). Cusp-like free-surface flows due to a submerged source or sink in
the presence of a flat or sloping bottom. J. Aust. Math. SOC. Ser. B. 26, 470-486.
Hocking, G. C. (1988). Critical withdrawal from a two-layer fluid through a Line sink.
Environmental Dynamics Rept. No. ED-88-240, Centre for Water Research, Univer-
sity of Western Australia.
Hocking, G. C., and Patterson, J. C. (1988). Quasi two-dimensional modelling of
reservoirs. Environmental Dynamics Rept. ED-88-283, Centre for Water Research,
University of Western Australia.
Hocking, G. C., Sherman, B. S., and Patterson, J. C. (1988). Algorithm for selective
withdrawal from stratified reservoir. J. Hydr. Div. ASCE 114,707-719.
Holloway, G. (1986). Eddies, waves, circulation and mixing: Statistical geofluid mechanics.
Ann. Rev. Fluid Mech. 18,91-147.
Holyer, J. Y.,and Huppert, H. E. (1980). Gravity currents entering a two-layer fluid. J.
Fluid Mech. 100,739-767.
Hopfinger, E. J. (1987). Turbulence in stratified fluids: A review. J. Geophys. Res. 92,
5287-5303.
Hopfinger, E. J., and Toly, J. A. (1976). Spatially decaying turbulence and its relation to
mixing across density interfaces. J . Fluid Mech. 78, 155-175.
Horn, W.,Mortimer, C. H., and Schwab, D. J. (1986). Wind-induced internal seiches in
Lake Zurich observed and modeled. Limnol. Oceanogr. 31, 1232-1254.
Howard, C. S. (1953). Density currents in Lake Mead. In “Proc. Intl. Hydr. Convention,”
ASCE. pp. 355-368. Minnesota.
Hsu, S. A. (1974). A dynamic roughness equation and its application to wind stress
determination at the air-sea interface. J . Phys. Oceanogr. 4, 116-120.
Hsu, S. A. (1986). A mechanism for the increase of wind stress (drag) coefficient with wind
speed over water surfaces: A parametric model. J. Phys. Oceanogr. 16, 144-150.
Physical Limnology 463
Huber, D. G. (1960). Irrotational motion of two fluid strata towards a Line sink. J. Eng.
Mech. Div.86, 71-86.
Huber, W. C., Harleman, D. R. F., and Ryan, P. J. (1972). Temperature prediction in
stratified reservoirs. J. Hydr.Div.ASCE 98, 645-666.
Hurley-Octavio, K. A., Jirka, G. H., and Harleman, D. R. F. (1977). Vertical heat
transport mechanisms in lakes and reservoirs. Tech. Rept. 227, Massachusetts Institute
of Technology.
Hussain, N. A., and Narang, B. S. (1984). Simplified analysis of air-bubble plumes in
moderately stratified environments. J. Heat Transfer 106,543-551.
Hutchinson, G. E. (1957). “A Treatise on Limnology. I. Geography, Physics and
Chemistry.” John Wiley & Sons, New York.
Hutter, K. (1984). “Hydrodynamics of Lakes.” Springer-Verlag, New York.
Hutter, K. (1986). Hydrodynamic modeling of lakes. In “Encyclopedia of Fluid Mechan-
ics,” pp. 897-998. Gulf Publishing, Houston, Texas.
Idso, S. B. (1973). On the concept of lake stability. Limnol. Oceunogr. 18,681-683.
Idso, S. B., and Cole, G. A. (1973). Studies on a Kentucky Knobs Lake: V. Some aspects
of the vertical transport of heat in the hypolimnion. J. Ecol. 61,413-420.
Imberger, J. (1970). Selective withdrawal from a stratified reservoir. Ph.D. thesis,
University of California, Berkeley.
Imberger, J. (1972). Two-dimensional sink flow of a stratified fluid contained in a duct. J.
Fluid Mech. 53, 329-349.
Imberger, J. (1974). Natural convection in a shallow cavity with differentially heated end
walls. Part 3: Experimental results. J. Fluid Mech. 65,247-260.
Imberger, J. (1976). Dynamics of a longitudinally stratified estuary. In “Proc. 15th Coastal
Eng. Conf.,” pp. 3108-3115. ASCE Hawaii.
Imberger, J. (1980). Selective withdrawal: A review. In “Proc. 2nd Intl. Symp. on Stratified
Flows,” Int. Ass. Hydraulic Research (IAHR) Trondheim.
Imberger, J. (1982). Reservoir dynamics modeling. In “Prediction in Water Quality” (E.
M. O’hughlin and P. Cullen eds.), pp. 223-248. Australian Academy of Science,
Canberra.
Imberger, J. (1985). The diurnal mixed layer. Limnol. Oceunogr. 30,737-770.
Imberger, J. (1987). Hydrodynamics of lakes. In “Proc. 12th Convention Australian Water
and Wastewater Assoc. ,” pp. 401-423. AWWA. Adelaide, South Australia.
Imberger, J. (1988). On the nature of turbulence in a stratified water body. Environmental
Dynamics Rept. No. ED-87-216, Centre for Water Research, University of Western
Australia.
Imberger, J., and Boashash, B. (1986). Application of the Wigner-Ville distribution to
temperature gradient microstructure: A new technique to study small-scale variations.
J. Phys. Oceunogr. 16, 1997-2012.
Imberger, J., and Fandry, C. (1975). Withdrawal of a stratified fluid from a vertical
two-dimensional duct. J. Fluid Mech. 70,321-332.
Imberger, J., and Fischer, H. B. (1970). Selective withdrawal from a stratified reservoir.
Water Pollution Control Res. Series 15040 EJZ 12/70, E.P.A. Water Quality
Office.
Imberger, J., and Hamblin, P. F. (1982). Dynamics of lakes, reservoirs and cooling ponds.
Ann. Rev. Fluid Mech. 14, 153-187.
Imberger, J., and Hebbert, R. H. B. (1980). Management of water quality in reservoirs.
Tech. Paper No. 49, Res. Proj. No. 74/70, Aust. Water Res. Council.
Imberger, J., and Monismith, S. G. (1986). A model for deepening due to upwelling when
W > 1. J. Fluid Mech. 171,432-439 (Appendix to Monismith (1986)).
464 Jorg Imberger and John C . Patterson
Imberger, J., and Parker, G. (1985). Mixed layer dynamics in a lake exposed to a spatially
variable wind field. Limnol. Oceanogr. 39,473-488.
Imberger, J., and Patterson, J. C. (1981). A dynamic reservoir simulation model-
DYRESM. In “Transport Models for Inland and Coastal Waters” (H. B. Fischer ed.),
pp. 310-361. Academic Press, New York.
Imberger, J., and Spigel, R. H. (1987). Circulation and mixing in Lake Rotongaio and Lake
Okaro under conditions of light to moderate winds: Preliminary results. N.Z. J . Marine
and Freshwater Res. 21,515-519.
Imberger, J., Thompson, R. 0. R. Y., and Fandry, C. (1976). Selective withdrawal from a
finite rectangular tank. J. Fluid Mech. 78,489-512.
Imberger, J., Patterson, J. C., Hebbert, R., and Loh, I. (1978). Dynamics of reservoirs of
medium size. J. Hydr. Diu. ASCE 104,725-743.
Imboden, D. M., and Emerson, S. (1977). Natural radon and phosphorus as limnologic
tracers: Horizontal and vertical eddy diffusion in Greifensee. Limnol. Oceanogr. 23,
77-90.
Imboden, D. M., and Joller, T. (1984). Turbulent mixing in the hypolimnion of
Baldeggersee (Switzerland) traced by natural radon-222. Limnol. Oceanogr. 29,
831-844.
Imboden, D. M., Weiss, R. F., Craig, H., Michel, R. L., and Goldman, C. R. (1977). Lake
Tahoe geochemical study. 1. Lake chemistry and tritium mixing study. Limnol.
Oceanogr. 22, 1039-1051.
Imboden, D. M., Lemmin, U., Joller, T., and Schurter, M. (1983). Mixing mechanisms in
lakes: Mechanisms and ecological relevance. Schweiz. Z. Hydrol. 45, 11-44.
Inaba, H., Seki, N., Fukusako, S., and Kanyama, K. (1981). Natural convective heat
transfer in a shallow rectangular cavity with different end temperatures. Numer. Heat
Transfer 4, 459-468.
Itsweire, E. C., Helland, K. N., and van Atta, C. W. (1986). The evolution of grid
generated turbulence in a stably stratified fluid. J. Fluid Mech. 162,299-338.
Ivetic, M., Iwasa, Y., and Inoue, K. (1986). Large eddy simulation of a shear driven flow in
a test reservoir. In “Proc. 2nd Intl. Conf. Hydrosoft 86,” (M. Radojkovic, C.
Maksimovic, and C. A. Brebbia, eds.) Southampton. pp. 209-216, Springer Verlag.
Ivey, G. N. (1987). The role of boundary mixing in the deep ocean. J . Geophys. Res. 92,
11873- 11878.
Ivey, G. N., and Blake, S. (1985). Axisymmetric withdrawal and inflow in a density-
stratified container. J. Fluid Mech. 161,115-137.
Ivey, G. N., and Boyce, F. M. (1982). Entrainment by bottom currents in Lake Erie.
Limnol. Oceanogr. 27, 1029-1038.
Ivey, G. N., and Corms, G. M.(1982). Boundary mixing in a stratified fluid. J . Fluid Mech.
121,1-26.
hey, G. N., and Hamblin, P. F. (1989). Convection near the temperature of maximum
density for high Rayleigh number, low aspect ratio, rectangular cavities. J . Heat
Transfer 111,100-105.
Ivey, G. N., and Nokes, R. I. (1989). Vertical mixing due to the breaking of initial waves
on sloping boundaries. J. Fluid Mech. 204,479-500.
Ivey, G. N., and Patterson, J. C. (1984). A model of the vertical mixing in Lake Erie in
summer. Limnol. Oceanogr. 29,553-563.
Jain, S. C. (1982). Buoyancy-driven circulation in free-surface channels. J. Fluid Mech. U2,
1-12.
Janowitz, G. S. (1986). A surface density and wind-driven model of the thermocline. J.
Geophys. Res. 91,5111-5118.
Physical Limnology 465
Jassby, A., and Powell, T. (1975). Vertical patterns of eddy diffusion during stratification in
Castle Lake, California. Limnol. Oceanogr. 20, 530-543.
Jirka, G. H., and Arita, M. (1987). Density currents or density wedges: Boundary-layer
influence and control methods. 1. Fluid Mech. 177, 187-206.
Jirka, G. H., and Katavola, D. S. (1979). Supercritical withdrawal from two-layered fluid
systems. Part 2: Three-dimensional flow into round intake. J. Hydr. Res. 17,
53-62.
Johnson, T. R., Ellis, C. R., Farrell, G. J., and Stefan, H. G. (1987a). Negatively buoyant
flow in a diverging channel. I: Flow regimes. 1. Hydr. Eng. 1l3,716-730.
Johnson, T. R., Ellis, C. R., Farrell, G. J., and Stefan, H. G. (1987b). Negatively buoyant
flow in a diverging channel. 11: 3-D flow field descriptions. J. Hydr. Eng. lW,731-742.
Johnson, T. R., Ellis, C. R., and Stefan, H. G. (1988). Negatively buoyant flow in a
diverging channel: IV: Entrainment and dilution. J . Hydr. Eng. lW,437-456.
Jokela, J. B., and Patterson, J. C. (1985). Quasi-two-dimensional modelling of reservoir
inflow. In “Proc. 21st IAHR Intl. Congr.,” Vol. 2, pp. 317-322. Melbourne,
Institution of Engineers, Aust. Natl. Conference Publication No. 85/13.
Kamial, J. C., Wyngaard, J. C., Haugen, D. A., Cote, 0. R., and Izumi, Y. (1976).
Turbulence structure in the convective boundary layer. J. A m o s . Sci. 33,2152-2169.
Kantha, L. H., Phillips, 0. M., and &ad, R. S. (1977). On turbulent entrainment at a
stable density interface. J. Fluid Mech. 79, 753-768.
Kao, T. W. (1970). Free streamline theory for inviscid stratified flow into a line sink. Phys.
Fluids W, 558-564.
Kao, T. W. (1976). Selective withdrawal criteria of stratified fluids. J. Hydr. Diu. A X E
102, 717-729.
Kao, T . W., Pao. H. P., and Wei, S. N. (1974). Dynamics of establishment of selective
withdrawal of a stratified fluid from a line sink. Part 2: Experiment. 1. Fluid Mech. 65,
689-710.
Kao, T. W., Park, C., and Pao, H. P. (1978). Inflows, density currents and fronts. Phys.
Fluids 21, 1912-1922.
Kato, H., and Phillips, 0. M. (1969). On the penetration of a turbulent layer into stratified
fluid. J . Fluid Mech. 37,643-655.
Keijman, J. Q. (1974). The estimation of the energy balance of a lake from simple weather
data. Boundary Layer Met. 7,399-407.
Keller, W. C., Plant, W. J., and Weissman, D. E. (1985). The dependence of X band
microwave sea return on atmospheric stability and sea state. 1. Geophys. Res. 90,
1019-1029.
Keulegan, G. G., and Brame, V. (1960). Fourteenth progress report on model laws for
density currents mixing effect of wind induced waves. NBS Project No. 0603-11-06431,
U.S. Dept. of Commerce.
Khomskis, W. R. (1969). “Dynamics and Thermics of Small Lakes” Vilnyus: Mintis (in
Russian).
Khomskis, W. R., and Filatova, T. N. (1972). Principles of typology of stratified lakes in
relation to vertical exchange. Verhandlungen der Internationalen Vereinigung fur
Theoretische und Angewandte Limnologie 18, 528-536.
Kielmann, J., and Simons, T. J. (1984). Some aspects of baroclinic circulation models. In
“Hydrodynamics of Lakes” (K. Hutter ed.), pp. 235-285. Springer-Verlag, New York.
Kim, B. R., Higgins, J. M., and Bruggink, D. J. (1983). Reservoir circulation patterns and
water quality. J . Enuiron. Eng. 109, 1284-1294.
Kirk, J. T. 0. (1983). “Light and Photosynthesis in Aquatic Ecosystems.” Cambridge
University Press, Cambridge.
466 Jorg Imberger and John C. Patterson
Kitaigorodskii, S. A. (1968). On the calculation of the aerodynamic roughness of the sea
surface. Izu. A m o s . Oceanic Phys. 4, 498-502.
Kitaigorodskii, S. A., and Zaslavskii, M. M. (1974). A dynamical analysis of the drag
conditions at the sea surface. Boundary Layer Met. 6 , 53-61.
Kitaigorodskii, S. A., Kutznetsov, 0. A., and Panin, G. N. (1973). Coefficients of drag,
sensible heat and evaporation in the atmosphere over the surface of the sea. Izu.
A m o s . Oceanic Phys. 9, 1135-1141.
Kitaigorodskii, S. A., Donelan, M. A., Lumley, J. A., and Terray, E. A. (1983).
Wave-turbulence interactions in the upper ocean. Part 11: Statistical characteristics of
wave and turbulent components of the random velocity field in the marine surface
layer. J . Phys. Oceanogr. l3(11), 1988-1999.
Kobus, H. E. (1968). Analysis of the flow produced by air bubble systems. In “Proc. 11th
Coastal Eng. Conf.,” pp. 1016-1031. ASCE London.
Koh, R. C. Y. (1966a). Unsteady stratified flow into a sink. J. Hydr. Res. 4(2), 21-34.
Koh, R. C. Y. (1966b). Viscous stratified flow towards a sink. J. Fluid Mech. 24,
555-575.
Kondo, J. (1975). Air-sea bulk transfer coefficients in diabatic conditions. Boundary Layer
Met. 9, 91-112.
Kranenburg, C. (1979). Destratification of lakes using bubble columns. J. Hydr. Diu.
ASCE 105, 333-349.
Kranenburg, C. (1984). Wind-induced entrainment in a stably stratified fluid. J . Fluid
Mech. 145, 253-273.
Kranenburg, C. (1985). Mixed-layer deepening in lakes after wind set-up. J. Hydr. Diu.
ASCE. 111, 1279-1297.
Kraus, E. B. (1977). “Modeling and Prediction of the Upper Layers of the Ocean.”
Pergamon Press, New York.
Kraus, E. B., and Turner, J. S. (1967). A one-dimensional model of the seasonal
thermocline: 11. The general theory and its consequences. Tellus 14,98-106.
Kullenberg, G., Murthy, C. R., and Westerberg, H. (1974). Vertical mixing characteristics
in the thermocline and hypolimnion regions of Lake Ontario. (IFYGL). In “Proc. 17th
Conf. Great Lakes Res.,” Vol. 1, pp. 425-434. Intl. Assoc. Great Lake Res. Ann
Arbor.
Lake Biwa Research Institute (1987). Fundamental studies on the vertical transport of
materials in Lake Biwa. Rept. No. 86-A05.
Langmuir, I. (1938). Surface motion of water induced by wind. Science 87, 119-123.
Large, W. G., and Pond, S. (1981). Open ocean momentum flux measurements in
moderate to strong winds. J. Phys. Oceanogr. 11, 324-336.
Lawrence, G. A., and Imberger, J. (1979). Selective withdrawal through a point sink in a
continuously stratified fluid with a pycnocline. Environmental Dynamics Rept. No.
ED-79-002, Centre for Water Research, University of Western Australia.
Lawrence, G. A., Lasheras, J. C., and Browand, F. K. (1987). Shear instabilities in
stratified flow. Preprint from “3rd Intl. Symp. on Stratified Flows,” Vol. 1. Pasadena,
California.
Leibovich, S. (1977a). Convective instability of stably stratified water in the ocean. J. Fluid
Mech. 82, 561-585.
Leibovich, S. (1977b). On the evolution of the system of wind drift currents and Langmuir
circulations in the ocean. Part 1: Theory and the averaged current. 1. Fluid Mech. 79,
715-743.
Leibovich, S. (1980). On wave-current interaction theories of Langmuir circulation. J. Fluid
Mech. 99, 715-724.
Physical Limnology 467
Leibovich, S. (1983). The form and dynamics of Langmuir circulations. Ann. Rev. Fluid
Mech. 15,391-427.
Leibovich, S . , and Paolucci, S. (1980). The Langmuir circulation instability as a mixing
mechanism in the upper ocean. J . Phys. Oceanogr. 10, 186-207.
Leibovich, S., and Radhakrishnan, K. (1977). On the evolution of the system of wind drift
currents and Langmuir circulations in the ocean. Part 2: Structure of the Langmuir
vortices. J. Fluid Mech. 80, 481-507.
Lerman, A. (1971). Time to chemical steady-state in lakes and ocean. A d v . in Chem. Ser.
106, 30-76.
Lerman, A., and Stiller, M. (1%9). Vertical eddy diffusion in Lake Tiberias.
Verhandlungen der Internationalen Vereinigung fur Theoretische und Angewandte
Limnologie 17,323-333.
Lewis, W. M. (1973). The thermal regime of Lake Lanao (Philipines) and its theoretical
implication for tropical lakes. Limnol. Oceanogr. 18, 200-217.
Lewis, E. L., and Perkin, R. G. (1982). Seasonal mixing processes in an Arctic fjord
system. J. Phys. Oceanogr. 12, 74-83.
Lighthill, M. J. (1969). Dynamic response of the Indian Ocean to the southwest monsoon.
Phil. Trans. R. SOC.London Ser. A . 26J, 45-92.
Lin, J. T., and Pao, Y. H. (1979). Wakes in stratified fluids. Ann. Rev. Fluid Mech. 11,
317-338.
Linden, P. F. (1975). The deepening of a mixed layer in a stratified fluid. J. Fluid Mech. 71,
385-405.
Linden, P. F., and Simpson, J. E. (1986). Gravity-driven flows in a turbulent fluid. J. Fluid
Mech. 172, 481-497.
Lo, A. K., and McBean, G. A. (1978). On the relative errors in methods of flux
calculations. J. Appl. Met. 17, 1704-1711.
Loftler, H. (1968). Tropical high-mountain lakes. Their distribution, ecology and zoo-
geographical importance. Colloquium Geographicum 9 , 57-76.
Long, R. R. (1956). Sources and sinks at the axis of a rotating liquid. Quart. J . Mech. and
Appl. Math. 9 , 385-393.
Lueck, R. G., and Osborn, T. R. (1985). Turbulence measurements in a submarine canyon.
Continental Shelf Res. 4, 681-698.
Luketina, D. A., and Imberger, J. (1987). Characteristics of a surface buoyant jet. J .
Geophys. Res. 92, 5435-5447.
Luketina, D. A,, and Imberger, J. (1988). Turbulence and entrainment in a buoyant
surface plume. J. Geophys. Res. (in press).
MacIntyre, S., and Melack, J. M. (1982). Meromixis in an equatorial African soda lake.
Limnol. Oceanogr. 27, 595-609.
Mahony, J. J., and Pritchard, W. G. (1977). Withdrawal from a reservoir of stratified fluid.
Rept. No. 79, Fluid Mech. Res. Institute, University of Essex.
Mahrt, L., and Lenschow, D. H. (1976). Growth dynamics of the convectively mixed layer.
J . A m o s . Sci. 33, 41-51.
Manins, P. C. (1976). Intrusion into a stratified fluid. J. Fluid Mech. 74, 547-560.
Markofsky, M., and Harleman, D. R. F. (1973). Prediction of water quality in stratified
reservoirs. J. Hydr. Div. ASCE 99, 729-745.
Marmoush, Y. R., Smith, A. A., and Hamblin, P. F. (1984). Pilot experiments on thermal
bar in lock exchange flow. J. Energy Eng. 110,215-227.
Marti, V. D. E., and Imboden, D. M. (1986). Thermische energieflusse an der
wasseroberflache: beispiel sempachersee. Schweiz. 2. Hydrol. 48, 196-229.
Maxworthy, T. (1972). Experimental and theoretical studies of horizontal jets in a
468 Jorg Zmberger and John C. Patterson
stratified fluid. In “IAHR Intl. Symp. on Stratified Flows,” Novosibirsk. IAHR. pp.
611-619.
Maxworthy, T. (1980). On the formation of nonlinear internal waves from the gravitational
collapse of mixed regions in two and three dimensions. J . Fluid Mech. 96, 47-72.
Maxworthy, T., and Monismith, S. G. (1988). Differential mixing in a stratified fluid. J.
Fluid Mech. 189,571-598.
McBean, G. A., and Paterson, R. D. (1975). Variations of the turbulent fluxes of
momentum, heat and moisture over Lake Ontario. J. Phys. Oceanogr. 5, 523-531.
McDougall, T. J. (1978). Bubble plumes in stratified environments. J. Fluid Mech. 85,
655-672.
McEwan, A. D., and Baines, P. G. (1974). Shear fronts and an experimental stratified
shear flow. J . Fluid Mech. 63, 257-272.
Mellor, G. L., and Durbin, P. A. (1975). The structure and dynamics of the ocean surface
mixed layer. J. Phys. Oceanogr. 5,718-728.
Mellor, G. L., and Yamada, T. (1982). Development of a turbulence closure model for
geophysical fluid problems. Rev. Geophys. Space Phys. 29,851-875.
Melville, W. K. (1977). Wind stress and roughness length over breaking waves. J. Phys.
Oceanogr. 7,702-710.
Milgram, J. H. (1983). Mean flow in round bubble plumes. J. Fluid Mech. l33,345-376.
Miyake, M., Donelan, M., McBean, G., Paulson, C., Badgley, F.,and Leavitt, E. (1970).
Comparison of turbulent fluxes over water determined by profile and eddy correlation
techniques. Quart. J . Roy. Met. SOC. 96, 132-137.
Monin, A. S., and Obukhov, A. M. (1954). Basic laws of turbulent mixing in the
atmosphere near the ground. Jr. Akad. Nauk SSSR Geofiz. Znst. 24(151), 163-187.
Monismith, S. G. (1985). Wind-forced motions in stratified lakes and their effect on
mixed-layer shear. Limnol. Oceanogr. 30,771-783.
Monismith, S . G. (1986). An experimental study of the upwelling response of stratified
reservoirs to surface shear stress. J. Fluid Mech. 171,407-439.
Monismith, S. G. (1987). The modal response of reservoirs to wind stress. J . Hydr. Diu.
ASCE 1l3,1290-1326.
Monismith, S. G., and Imberger, J. (1988). Horizontal convection in the sidearm of a small
reservoir. Environmental Dynamics Rept. ED-86-161, Centre for Water Research,
University of Western Australia.
Monismith, S . G., and Maxworthy, T. (1988). Selective withdrawal and spin-up of a
rotating, stratified fluid. J . Fluid Mech. 199,377-401.
Monismith, S. G., Imberger, J., and Billi, G. (1988). Shear-waves and unsteady selective
withdrawal. J . Hydr. Diu. ASCE 114,1134-1152.
Mortimer, C. H. (1974). Lake hydrodynamics. 1.V .L . u),124.
Mortimer, C. H. (1987). Fifty years of physical investigations and related limnological
studies on Lake Erie, 1928-1977. J . Great Lakes Res. l3,407-435.
Mulhearn, P. J. (1981). On the formation of a stably stratified internal boundary-layer by
advection of warm air over a cooler sea. Boundary Layer Met. 21,247-254.
Muller, P., and Garwood, R. W. (1988). Mixed layer dynamics: Progress and new
directions. EOS 69, 2-4, 10-12.
Muller, P., and Henderson, D. (1987). Dynamics of the oceanic surface mixed layer. In
“Proc. Aha Huliko’a Hawaiian Winter Workshop,” Hawaii Institute of Geophysics
Special Publication, Honolulu.
Muller, P., and Pujalet, R. (1984). Internal gravity waves and small-scale turbulence. In
“Proc. Aha Huliko’a Hawaiian Winter Workshop,” Hawaii Institute of Geophysics
Special Publication, Honolulu.
Physical Limnology 469
Munk, W., and Anderson, E. R. (1948). Notes on a theory of the thermocline. J. Marine
Res. 7 , 276-295.
Murota, A., and Hirata, T. (1979). Breaking of internal gravity wave on slope. Tech. Rep&.
of Osaka University 29(1487), 255-264.
Murota, A., and Michioku, K. (1986a). Analysis on selective withdrawal from three-layered
stratified systems and its practical application. J. Hydrosci. Hydr. Eng. 4, 31-50.
Murota, A., and Michioku, K. (1986b). Characteristics of internal fluctuations in thermally
stratified fields induced by combination of mechanical and thermal stirring. Proc. JSCE
375(II-6), 171-179.
Murota, A., Hirata, T., and Michioku, K. (1980). Breaking of internal gravity wave on
slope: 11. Tech. Rep&. of Osaka University 30(1543), 255-261.
Murota, A., Michioku, K., and Kuchida, M. (1988). Numerical analysis on vertical mixing
phenomena induced by composite stirring. Proc. JSCE 11-9, 67-76.
Myrup, L. 0.. Powell, T. M., Godden, D. A., and Goldman, C. R. (1979). Climatological
estimate of the average monthly energy and water budgets of Lake Tahoe, California-
Nevada. Water Resources Res. 15, 1499-1508.
Narimousa, A., and Fernando, H. J. S. (1987). On the sheared-density interface of an
entraining stratified fluid. J . Fluid Mech. 174, 1-22.
Narimousa, S., Long, R. R., and Kitaigorodskii, S . A. (1986). Entrainment due to
turbulent shear flow at the interface of a stably stratified fluid. Tellus %A,
7687.
Newman, F. C. (1976). Temperature steps in Lake Kivu: A bottom heated saline lake. J.
Phys. Oceanogr. 6 , 157-163.
Niiler, P. P. (1975). Deepening of the wind-mixed layer. J . Marine Res. 33,405-422.
Niiler, P. P., and Kraus, E. B. (1977). One-dimensional models of the upper ocean. In
“Modeling and Prediction of the Upper Layers of the Ocean” (E.B. Kraus ed.), pp.
143-172, Pergamon Press, New York.
Nishida, S., and Yoshida, S . (1984). Stability of a two-layer shear flow. Theoretical and
App. Mech. 32, 35-45.
Nokes, R. I. (1988). On the entrainment rate across a density interface. 1. Fluid Mech. 188,
185-204.
O’Donnell, J., and Garvine, R. W.(1983). A time dependent, two-layer frontal model of
buoyant plume dynamics. Tellus 35A, 73-80.
Ookubo, K., and Muramoto, Y. (1981). Coriolis’ effects on density CurrentsExperiments
on the rotating exchange flow. Bulletin of the Disaster Prevention Research Institute
24B-2,339-365.
Ookubo, K., Muramoto, Y., Oonishi, Y., and Kumagai, M. (1984). Laboratory experi-
ments on thermally induced currents in Lake Biwa. Bulletin of the Dbaster Prevention
Research Institute 34/2(304), 19-54.
Orlob, G. T., and Selna, L. G. (1970). Temperature variations in deep reservoirs. 1. Hydr.
Div. ASCE 96.391-410.
Panofsky, H. A. (1983). Determination of stress from wind and temperature measure-
ments. Quart. J. Roy. Meteor. SOC.89, 85-94.
Pao, H. P., and Kao, T. W. (1969). Sources and sinks at the axis of a viscous rotating fluid.
Phys. Fluiah 12, 1536-1546.
Pao, H. P., and Kao, T. W. (1974). Dynamics of establishment of selective withdrawal of a
stratified fluid from a line sink. Part I: Theory. 1. Fluid Mech. 65, 657-688.
Pao, H. P., and Shih, T. H. (1973). Selective withdrawal and blocking wave in rotating
fluids. J . Fluid Mech. 57,459-480.
Parker, G. J., and Imberger, J. (1986). Differential mixed-layer deepening in lakes and
470 Jorg Imberger and John C . Patterson
reservoirs. In “Limnology in Australia” (P. de Deckker and W. D. Williams eds.), pp.
63-92. W. Junk Publishers, Melbourne, Victoria.
Patalas, K. (1960). Mieszanie wody jako czynnik okreslajacy intensywnose krazenia materii
w roznych morfologicznie jeziorach okolic Wegorzewa. Roczniki Nauk Rolniczych.
B77,223-242.
Patalas, K. (1961). Wind und morphologiebedingte wasserwegunstypen als bestimmender
Faktor fur die Intersitat des Stoffkreislaufes in nordpolnischen Seen. Verhandlungen
der Internationalen Vereinigung fur Theoretische und Angewandte Limnologie. 14,
59-64.
Patterson, J. C. (1984). Unsteady natural convection in a cavity with internal heating and
cooling. J. Fluid Mech. 140, 135-151.
Patterson, J. C., and Hamblin, P. F. (1988). Thermal simulation of a lake with winter ice
cover. Limnol. Oceanogr. 33, 323-338.
Patterson, J., and Imberger, J. (1980). Unsteady natural convection in a rectangular cavity.
J , Fluid Mech. 100,65-86.
Patterson, J. C., and Imberger, J. (1989). Simulation of bubble plume destratification
systems in reservoirs. Aquatic Sciences 51,3-18.
Patterson, J. C., Hamblin, P. F., and Imberger, J. (1984). Classification and dynamic
simulation of the vertical density structure of lakes. Limnol. Oceanogr. 29, 845-861.
Patterson, J. C., Allanson, B. R., and hey, G. N. (1985). A dissolved oxygen budget
model for Lake Erie in summer. Freshwater Biology 15,683-694.
Paulson, C. A. (1970). The mathematical representation of wind speed and temperature
profiles in the unstable atmospheric surface layer. J . Appl. Met. 9, 857-861.
Peterson, E. W. (1969). Modification of mean flow and turbulent energy by a change in
surface roughness under conditions of neutral stability. Quart. J . Roy. Met SOC.95,
561-575.
Phillips, 0 . M., Shyu, J. H., and Salmun, H. (1986). An experiment on boundary mixing:
Mean circulation and transport rates. J. Fluid Mech. 173,473-499.
Pollard, R. T. (1977). Observations and theories of Langmuir circulations and their role in
near suface mixing. In “A Voyage of Discovery: George Deacon 70th Anniversary
Volume” (M. Angel ed.), pp. 235-251. Pergamon Press, New York.
Pollard, R. T., Rhines, P. B., and Thompson, R. 0. R. Y.(1973). The deepening of the
wind mixed layer. Geophys. Fluid Dyn. 3, 381-404.
Pond, S.,Fissel, D. B., and Paulson, C. A. (1974). A note on bulk aerodynamic coefficients
for sensible heat and moisture fluxes. Boundary Layer Met. 6 , 333-339.
Price, J. F., Mooers, C. N. K., and Van Leer, J. C. (1978). Observations and simulation of
storm-induced mixed-layer deepening. J. Phys. Oceanogr. 8 , 582-599.
Raphael, J. M. (1962). Prediction of temperature in rivers and reservoirs. J . Power Diu. 88,
157- 182.
Rayner, K. N. (1981). Diurnal energetics of a reservoir surface layer. Environmental
Dynamics Rept. ED-80-005,Centre for Water Research, University of Western
Australia.
Robarts, R. D., and Ward, P. R. B. (1978). Vertical diffusion and nutrients transport in a
tropical lake (Lake McIlwaine, Rhodesia). Hydrobiologia 59, 213-221.
Robarts, R. D., Ashton, P. J., Thornton, J. A., Taussig, H. J., and Sephton, L. M. (1982).
Overturn in a hypertonic, warm, monomictic impoundment (Hartbeespoort Dam,
South Africa). Hydrobiologia 97, 209-224.
Roberts, P. J. W., and Matthews, P. R. (1987). Behavior of low buoyancy jets in a linearly
stratified fluid. 1. Hydr. Res. 25, 503-519.
Rodgers, G. IG. (1971). Field investigation of the thermal bar in Lake Ontario: Precision
Physical Limnology 47 1
temperature measurements. In “Proc. 14th Conf. Great Lakes Res.,” pp. 618-624.
Intl. Assoc. Great Lakes Res. Ann Arbor.
Rodi, W. (1982). “HMT, Vol. 6: Turbulent buoyant jets and plumes.” Pergamon Press,
Sydney.
Rodi, W. (1987). Examples of calculation methods for flow and mixing in stratified fluids. J.
Geophys. Res. 92(C5), 5305-5328.
Rohr, J. J., Itsweire, E. C., Helland, K. N., and Van Atta, C. W.(1988). An investigation
of the growth of turbulences in a uniform-mean-shear flow. 1. Fluid Mech. 187,
1-33.
Sadhuram, Y., Vethamony, P., Suryanarayana, A., Swamy, G. N., and Sastry, J. S. (19b8).
Heat energy exchange over a large water body under stable atmospheric conditions.
Boundary Layer Met. 44, 171-180.
Sahlberg, J. (1983). A hydrodynamical model for calculating the vertical temperature
profile in lakes during cooling. Nordic Hydrology 239-254.
Sanderson, B., Perry, K., and Pederson, T. (1986). Vertical diffusion in meromictic Powell
Lake, British Columbia. J. Geophys. Res. 91, 7647-7655.
Sargent, R. E., and Jirka, G. H. (1982). A comparative study of density currents and
density wedges (with application to intermediate field dynamics of ocean thermal
energy conversion plants). Contract No. ET-78-S-02-4683, U.S.Dept. of Energy.
Savage, S. B., and Brimberg, J. (1975). Analysis of plunging phenomena in water
resources. J. Hydr. Res. W, 187-204.
Schertzer, W. M., Saylor, J. H., Boyce, F. M., Robertson, D. G., and Rosa, F. (1987).
Seasonal thermal cycle of Lake Erie. J. Great Lakes Res. 13,468-486.
Scott, C. F., and Imberger, J. (1988). Three-dimensional estuary circulation and class-
ification. In “Proc. Intl. Conf. Phys. of Shallow Estuaries and Bays, (R. Cheng, ed.)
Springer-Verlag. Pacific Grove, California. Lecture Notes on Coastal and Estuarine
Studies.
Scott, J. T., Myer, G. E., Stewart, R., and Walther, E. G. (1969). On the mechanism of
Langmuir circulations and their role in epilimnion mixing. Limnol. Oceanogr. 14,
493-503.
Shay, T. J., and Gregg, M. C. (1984). Turbulence in an oceanic convective mixed layer.
Nature 310, 282-285.
Shay, T. J., and Gregg, M. C. (1986). Convectively driven turbulent mixing in the upper
ocean. J. Phys. Oceanogr. 16, 1777-1798.
Shay, T. J., and Imberger, J. (1988). Mixing in a strongly stratified low-energy lake.
Environmental Dynamics Rept. ED-88-255, Centre for Water Research, University of
Western Australia.
Sherman, F. S., Imberger, J., and Corcos, G. M. (1978). Turbulence and mixing in stably
stratified waters. Ann. Rev. Fluid Mech. 10, 267-288.
Shih, H. H., and Pao, H. S. (1971). Selective withdrawal in rotating fluids. J. Fluid Mech.
49,509-527.
Shih, T. H., Lumley, J. L., and Janicka, J. (1987). Second-order modeling of a
variable-density mixing layer. J. Fluid Mech. 180, 93-116.
Silvester, R. (1979). An experimental study of end wall effects on selective withdrawal of
linearly stratified liquid from a reservoir. Ph.D. thesis, University of Western
Australia.
Simpkins, P. G., and Dudderar, J. D. (1981). Convection in rectangular cavities with
differentially heated end walls. J. Fluid Mech. 110, 433-456.
Simpson, J. E. (1982). Gravity currents in the laboratory, atmosphere, and ocean. Ann.
Rev. Fluid Mech. 14, 213-234.
472 Jorg Imberger and John C. Patterson
Smith, S. D. (1980). Wind stress and heat flux over the ocean in gale force winds. J. Phys.
Oceanogr. 5 , 709-726.
Smith, S. D., and Anderson, R.J. (1984). Spectra of humidity, temperature and wind over
the sea at Sable Island, Nova Scotia. J . Geophys. Res. 89,2029-2040.
Smith, S. D., and Banke, E. G. (1975). Variation of the sea surface drag coefficient with
wind speed. Quart. J. Roy. Met. SOC.101, 665-673.
Smith, J., Pinkel, R., and Weller, R. A. (1987). Velocity structure in the mixed layer
during MILDEX. J . Phys. Oceanogr. 17,425-439.
Smyth, N. F., and Holloway, P. E. (1988). Hydraulic jump and undular bore formation on
a shelf break. J . Phys. Oceanogr. 18,947-962.
Sorbjan, Z.(1986). On similarity in the atmospheric boundary layer. Boundary Layer Met.
34, 377-397.
Spain, J. D., Wernert, G. M., and Hubbard, D. W. (1976). The structure of the spring
thermal bar in Lake Superior. J. Great Lakes Res. 2, 296306.
Spalding, D. B., and Svensson, U. (1977). The development and erosion of the
thermocline. In “Proc. Seminar Intl. Centre for Heat and Mass Transfer,” Hemisphere
Publ. Corp. Washington D.C. Vol. 1, pp. 113-122.
Spigel, R. H. (1978). Wind mixing in lakes. Ph.D. thesis, University of California,
Berkeley.
Spigel, R. H. (1980). Coupling of internal wave motion with entrainment at the density
interface of a two-layer lake. J . Phys. Oceanogr. 10, 144-155.
Spigel, R. H., and Farrant, B. (1984). Selective withdrawal through a point sink and
pycnocline formation in a linearly stratified flow. 1. Hydr. Res. 22, 35-51.
Spigel, R. H.,and Imberger, J. (1980). The classification of mixed layer dynamics in lakes
of small to medium size. J . Phys. Oceanogr. 19, 1104-1121.
Spigel, R. H., Imberger, J., and Rayner, K. N. (1986). Modeling the diurnal mixed layer.
Limnol. Oceanogr. 31,533-556.
Stacey, M. W., and Bowen, A. J. (1988). The vertical structure of density and turbidity
currents: Theory and observations. J. Geophys. Res. 93,3528-3542.
Stefan, H.G.,and Ford, D. E. (1975). Temperature dynamics in dimictic lakes. J . Hydr.
Diu. A X E 101,97-114.
Stefan, H. G., Cardoni, J. J., Schiebe, F. R., and Cooper, C. M. (1983). Model of light
penetration in a turbid lake. Water Resources Res. 19, 109-120.
Steinhorn, I., and Assaf, G. (1980). The physical structure of the Dead Sea water column:
1975-1977. In “Hypersaline Brines and Evaporitic Environments” (A. Nissenbaum
ed.), pp. 145-153. Elsevier, Amsterdam.
Steinhorn, I., and Gat, J. R. (1983). The Dead Sea. Scienttjic American. 249(4), 102-109.
Stewart, R. W. (1974). The air-sea momentum exchange. Boundary Layer Met. 6 ,
151-167.
Stocker, T., and Hutter, K. (1986). One-dimensional models for topographic Rossby waves
in elongated basins on the f-plane. J . Fluid Mech. 170,435-459.
Straskraba, M. (1980). The effects of physical variables on freshwater production: Analyses
based on models. In “The Functioning of Freshwater Ecosystems” (E. D. Le Cren and
R. H. McConnell eds.), pp. 13-84. Cambridge University Press, England.
Strub, P. T. (1983).The response of a small lake to atmospheric forcing during fall cooling.
Ph.D. thesis, University of California.
Strub, P. T., and Powell, T. M. (1986). Wind-driven surface transport in stratified closed
basins: Direct versus residual circulations. J. Geophys. Res. 91, 8497-8508.
Strub, P. T., and Powell, T. M. (1987). The exchange coefficients for latent and sensible
heat flux over lakes: Dependence upon atmospheric stability. Boundary Layer Met. 40,
349-361.
Physical Limnology 473
Stull, R. (1976). The energetics of entrainment across a density interface. J . Amos. Sci. 33,
1260- 1267.
Sturm, T. W. (1981). Laminar gravitational convection of heat in dead-end channels. J .
Fluid Mech. 110, 97-113.
Sun, T. Y., and Faeth, G. M. (1986). Structure of turbulent bubbly jets-1. Methods and
centerline properties. Inrl. J. Mukiphare Flow 12, 99-114.
Sundaram, T. R., and Rehm, R. G. (1973). The seasonal thermal structure of deep
temperate lakes. Tellus 25, 157-168.
Svensson, U. (1978). A mathematical model of the seasonal thermocline. Rept. No. 1002,
Dept. of Water Resources Eng., University of Lund, Sweden.
Sverdrup, H. U. (1945). “Oceanography for Meteorologists.” George Allen and Unwin,
London.
Swinbank, W. C. (1963). Long-wave radiation from clear skies. Quurr. J . Roy. Met. SOC.
89,339-348.
Tacke, K. H., Schubert, H. G., Weber, D. J., and Schwerdtfeger, K. (1985). Characteris-
tics of round vertical gas bubble jets. Metallurgical Trans. B. 16, 263-275.
Taylor, G. I. (1954). The dispersion of matter in turbulent flow through a pipe. Proc. Roy.
SOC. London Ser. A . 223,446-468; (1960). Sci. Pap. 2 , 466-488.
Taylor, P. A. (1970). A model of airflow above changes in surface heat flux, temperature
and roughness for neutral and unstable conditions. Boundary Layer Mer. 1,474-497.
Taylor, M., and Aquise, E. (1984). A climatological energy budget of Lake Titicaca
(Peru/Bolivia). I.V.L . 22, 1246-1251.
Tennekes, H., and Lumley, J. L. (1972). “A First Course in Turbulence,” MIT Press,
Cambridge, Massachusetts.
Tennessee Valley Authority (1972). Heat and Mass Transfer Between a Water Surface and
the Atmosphere. Lab. Rept. No. 14, Noms, Tennessee.
Thomas, J. H. (1975). A theory of steady wind-driven currents in shallow water with
variable eddy viscosity. 1. Phys. Oceanogr. 5 , 136-142.
Thompson, R. 0. R. Y., and Imberger, J. (1980). Response of a numerical model of a
stratified lake to wind stress. In “Proc. 2nd IAHR Symp. on Stratified Flows,
Trondheim.” (T. Carstens and T. McClimans eds.), pp. 562-570. IAHR.
Thorpe, S. A. (1969). Experiments on stability of stratified shear flow. Radio Sci. 4,
1327- 1331.
Thorpe, S . A. (1973). Turbulence in stably stratified fluids: A review of laboratory
experiments. Boundary Layer Mer. 5.95-119.
Thorpe, S . A. (1974). Near-resonant forcing in a shallow two-layer fluid: A model for the
internal surge in Loch Ness? J . Fluid Mech. 63, 509-527.
Thorpe, S. A. (1978). The near-surface ocean mixing layer in stable heating conditions. J .
Geophys. Res. 8 , 2875-2885.
Thorpe, S. A. (1982). On the layers produced by rapidly oscillating a vertical grid in a
uniformly stratified fluid. J. Fluid Mech. W, 391-409.
Thorpe, S . A. (1987). Transitional phenomena and the development of turbulence in
stratified fluids: A review. J. Geophys. Res. 92, 5231-5248.
Thorpe, S. A., and Hall, A. J. (1982). Observations of the thermal structure of Langmuir
circulation. J. Fluid Mech. 114, 237-250.
Tolland, H. G. (1977). Destratification/aeration in reservoirs. Tech. Rept. TR 50. Water
Research Centre, England.
Trevisan, 0. V., and Bejan, A. (1986). Convection driven by the non-uniform absorption
of thermal radiation at the free surface of a stagnant pool. Numerical Hear Transfer 10,
483-506.
Trusturn, K. (1964). Rotating and stratified flow. J . Fluid Mech. 19, 415-432.
474 Jorg Zmberger and John C . Patterson
Tsihrintzis, V. A., and Alavian, V. (1987). A model for predicting the properties of
three-dimensional negatively buoyant starting plumes on a sloping surface. In “Proc.
3rd Intl. Symp. on Stratified Flows,” California Institute of Technology, Pasadena.
IAHR.
Tuck, E. O., and Vanden Broeck, J. M. (1984). A cusp-like free-surface flow due to a
submerged source or sink. J. Aust. Math. SOC. Ser. B 25, 443-450.
Tucker, W. A,, and Green, A. W. (1977). A time-dependent model of the lake averaged,
vertical temperature distribution of lakes. Limnol. Oceanogr. 22, 687-699.
Turner, J. S. (1985). Multicomponent convection. Ann. Rev. Fluid Mech. 17, 11-44.
Tzur, Y. (1973). One-dimensional diffusion equations for the vertical transport in an
oscillating stratified lake of varying cross-section. Tellus 25, 266-271.
UNESCO (1981). Tenth report of the joint panel on oceanographic tables and standards.
Sidney, B.C., Canada.
U.S. Army Coastal Engineering Research Center (1977). “Shore Protection Manual,” Vol.
1, 3.35-3.37. U.S. Govt. Printing Office,Washington.
U.S. Army Corps of Engineers (1982). CE-QUAL-R1: A numerical one-dimensional
model of reservoir water quality. Instr. Rept. No. E-82-1, Waterways Experiment
Station.
Vanden Broeck, J. M., and Keller, J. B. (1987). Free surface flow due to a sink. J. Fluid
Mech. 175, 109-117.
Venkatram, A. (1977). A model of internal boundary-layer development. Boundary Layer
Met. 11,419-437.
Venkatram, A. (1986). An examination of methods to estimate the height of the coastal
internal boundary layer. Boundary Layer Met. 36, 149-156.
Ventz, D. (1973). Die Einzugsgebietsgrosse, ein Geofaktor fur den Trophiczustand
stehender Gewasser. Forrschnitte der Wusserchemie und ihrer Grenzgebiete 14, 105-
118.
Veronis, G. (1967). Analogous behavior of homogeneous, rotating fluids and stratified,
nonrotating fluids. Tellus 19,620-634.
Walesh, S. G., and Monkmeyer, P. L. (1973). Bottom withdrawal of viscous stratified fluid.
J. Hydr. Div. ASCE 99, 1401-1418.
Wallace, B. C., and Wilkinson, D. L. (1988). Run-up of internal waves on a gentle slope in
a two-layered system. J. Fluid Mech. 191,419-442.
Ward, P. R. B. (1977). Diffusion in lake hypolimnia. In “Proc. 17th Congr. Intl. Assoc.
Hydr. Res.,” Vol. 2, pp. 103-110. Baden-Baden. IAHR.
Water Resources Engineers, Inc. (1969). Mathematical models for prediction of thermal
energy changes in impoundments. U.S. E.P.A. Water Pollution Control Res. Series,
16130 EXT, Contract 14-22-422. U.S. Govt. Printing Office, Washington, D.C.
Wedderburn, E. M. (1906). The temperature of the freshwater lochs of Scotland with
special reference to Loch Ness. Trans. Roy. SOC. Edinburgh 45, 407-490.
Wedderburn, E. M. (1911). The temperature seiche: Observations in the Madusee. Trans.
Roy. SOC. Edinburgh 47,619-642.
Wedderburn, E. M. (1912). Temperature observations in Loch Earn with a further
contribution to the hydrodynamical theory of temperature seiches. Trans. Roy. SOC.
Edinburgh 48,629-695.
Weinstock, J. (1984). Effect of gravity waves on turbulence decay in stratified fluids. J.
Fluid Mech. 140,11-26.
Weiss, W., Lehn, H., Munnich, K. O., and Fischer, K. H. (1979). On the deep-water
turnover of Lake Constance. Arch. Hydrobiol. 86, 405-422.
Weller, R. A., Dean, J. P., Marra, J., Price, J. F., Francis, E. A., and Boardman, D. C.
Physical Limnology 475

(1985). Three dimensional flow in the upper ocean. Science 227, 1552-1556.
Wetzel, R. G. (1975). “Limnology.” W. B. Saunders, Philadelphia, Pennsylvania.
Wieringa, J. (1974). Comparison of three methods for determining strong wind stress over
Lake Flevo. Boundary Layer Met. 7,3-19.
Wilkinson, D. L. (1972). Dynamics of contained oil slicks. J. Hydr. Diu. ASCE 98,
1013-1030.
Wilkinson, D. L. (1979). Two-dimensional bubble plumes. J. Hydr. Diu. ASCE 105,
139-154.
Wilkinson, D. L., and Wood, I. R. (1972). Some observations on the motion of the head of
a density current. J. Hydr. Res. 10, 305-324.
Willis, G. E., and Deardorf€, J. W. (1974). A laboratory model of the unstable planetary
boundary layer. J. Atmos. Sci. 31, 1297-1307.
Witten, A. J., and Thomas, J. H. (1976). Steady-wind driven currents in a large lake with
depth-dependent eddy viscosity. J. Phys. Oceanogr. 6, 85-92.
Wood, I. R. (1978). Selective withdrawal from two-layer fluid. 1. Hydr. Diu. ASCE 104,
1647- 1659.
Wright, J. C. (1961). The limnology of Canyon Ferry Reservoir. IV. The estimation of
primary production from physical limnological data. Limnol. Oceanogr. 6 , 330-337.
Wu, J. (1973). Wind-induced turbulent entrainment across a stable density interface. J.
Fluid Mech. 61, 275-287.
Wu, J. (1980). Wind-stress coefficientsover sea surface near neutral conditions-A revisit. J.
Phys. Oceanogr. 10,727-740.
Wu,J. (1985). Parameterization of wind-stress coefficientsover water surfaces. J. Geophys.
Res. 90,9069-9072.
Wucknitz, J. (1976). Determination of turbulent fluxes of momentum and sensible heat
from fluctuation measurements and the structure of wind field over waves above the
tropical Atlantic during ATEX. “Meteor” Forschungsergab 11, 25-50.
Wuest, A. (1987). Ursprung und grosse von mischungsprozessen im hypolimnion natur-
licher seen. Rept. No. 8350, Eidgenossischen Technischen Hoschschule, Zurich.
Wuest, A., Imboden, D. M., and Schurter, M. (1988). Origin and size of hypolimnic mixing
in Urnersee, the southern basin of Vienvaldstattersee (Lake Lucerne). Schweiz. Z .
Hydrol. 50.
Wunderlich, W. D., and Elder, R. A. (1969). Selective withdrawal from density stratified
reservoirs. Rept. No. 21, Water Resources Res. Advance.
Yih, C. S. (1965). “Dynamics of non-homogeneous fluids.” Macmillan, New York.
Yih, C. S. (1980). “Stratified Flows.’’ Academic Press, New York.
Yoshimura, S. (1936). A contribution to the knowledge of deep water temperature of
Japanese lakes. Japanese J. Astro. Geophys. 13, 61-120.
Zeman, O., and Tennekes, H. (1977). Parameterization of the turbulent energy budget at
the top of the daytime atmospheric boundary layer. J. Amos. Sci. 34, 111-123.

You might also like