Sie sind auf Seite 1von 12

989-1000 GW Watershed 03 12/1/03 4:53 PM Page 989

Where Does the Ground Water


in Small Watersheds Come From?
by Thomas C. Winter1, Donald O. Rosenberry1, and James W. LaBaugh2

Abstract
Surface water and ground water watersheds commonly do not coincide. This condition is particularly relevant to
understanding biogeochemical processes in small watersheds, where detailed accounting of water and solute fluxes
commonly are done. Ground water watersheds are not as easily defined as surface watersheds because (1) they are
not observable from land surface; (2) ground water flow systems of different magnitude can be superimposed on one
another; and (3) ground water divides may move in response to dynamic recharge and discharge conditions. Field
studies of relatively permeable terrain in Wisconsin, Minnesota, and Nebraska indicate that lakes and wetlands in
small watersheds located near the lower end of extensive ground water flow systems receive ground water inflow
from shallow flow systems that extend far beyond their surface watershed, and they may also receive ground water
inflow from deeper regional flow systems that pass at depth beneath local flow systems. Field studies of mountain-
ous terrain that have low-permeability deposits in New Hampshire and Costa Rica also indicate that surface water
bodies receive ground water inflow from sources beyond their local surface watersheds. Field studies of lakes and
wetlands in North Dakota, Nebraska, and Germany indicate that ground water divides move in response to changing
climate conditions, resulting in a variable source of ground water inflow to those surface water bodies.

Introduction water (Likens et al. 1967; Schindler et al. 1976; Hemond


Small watersheds have been the focus of research on 1980). In other cases, although contribution of water from
the hydrology and biogeochemistry of various types of ground water was small, such contribution delivered sub-
land-use landscapes, including agricultural, forested, urban, stantial solutes (Schwartz and Gallup 1978; Hurley et al.
and natural (Hewlett et al. 1969; Winter 1997; McCammon 1985). Subsequent examination of watershed processes in a
et al. 1998; Baedecker and Friedman 2000). Small water- variety of landscapes has shown surface water can receive
sheds commonly are selected for intensive research considerable water and solute contributions from ground
because they are believed to be areas that are well defined water (LaBaugh et al. 1995; Moore 1999; Holmes 2000).
by the watershed boundary. By being hydrologically well Thus, there is growing recognition that watershed process
defined, an area lends itself to detailed accounting of gains, studies should include examination of ground water
losses, and storage of water and solutes. (Goodrich and Woolhiser 1993).
The need to examine watershed processes in order to A watershed is the topographic demarcation that
understand flow and transport in surface waters has long defines a surface water drainage basin, and by doing so,
been recognized (Leopold 1974; Hynes 1975; Likens separates surface drainage basins from one another. Tradi-
1984). Some detailed studies of surface waters that tionally, the term “watershed” is used with respect to sur-
included an accounting of water and solutes were in basins face water. Ground water hydrologists use the term less
that had negligible contributions from or losses to ground often because they traditionally deal with aquifer systems.
However, ground water systems also have watersheds, at
flow-system divides. Although it would be convenient for
water- and chemical-budget studies of watersheds if ground
1U.S. Geological Survey, Box 25046, Denver Federal Center,
water divides underlay surface divides, they commonly do
Denver, CO 80225
2U.S. Geological Survey, Mail Stop 411 National Center, not. This is especially true for small watersheds. Further-
12201 Sunrise Valley Dr., Reston, VA 20192 more, as natural resource managers move increasingly
Published in 2003 by the National Ground Water Association. toward managing water supply, water quality, and other
Vol. 41, No. 7—GROUND WATER—Watersheds Issue 2003 (pages 989–1000) 989
989-1000 GW Watershed 03 12/1/03 4:53 PM Page 990

resources on the basis of surface watersheds (Sosin et al. defined by Tóth were applicable primarily to terrain having
1995), it is becoming increasingly important that the extent low-hydraulic conductivity (generally less than ~0.3
of ground water watersheds be considered in many of those m/day). Ground water flow systems in porous media having
management plans and practices. hydraulic conductivity greater than this are less likely to
The common assumption that ground water divides have water table mounds underlying topographic highs
underlie surface divides probably stems from the work of (Figure 1c), in which case the water table is more likely to
Hubbert (1940), whose classic diagram showed ground be a continuously sloping surface from one surface water
water moving from water table highs beneath uplands to body to the next.
contiguous lowlands (Figure 1a). Tóth (1962), based largely Ground water watersheds are not as easily defined as
on work in the Canadian prairies, extended this concept for surface watersheds because (1) they are not observable
much larger regions and showed further that flow systems of from land surface, (2) ground water flow systems of differ-
different magnitudes could overlie one another (Figure 1b). ent magnitude can be superimposed on one another, and (3)
Haitjema (1995) indicated that the type of flow systems ground water divides may move in response to dynamic

A Land surface
Water table

Direction of ground water flow

B
Line separating local flow
METERS
systems from each other and
80 from regional flow systems
Land surface Wetland
60

50 Water table

40

30

20
Aquifer
10
Direction of ground water flow
0
0 2000 4000 6000 8000 10,000 12,000 14,000 16,000
METERS

C
METERS
60 Land surface

50 Wetland
Water table
40

30
Direction of ground water flow
20

10

0
0 2000 4000 6000 8000 10,000 12,000 14,000 16,000
METERS

Figure 1. Hydrologic sections showing ground water flow systems in hypothetical settings: (a) ground water flow from water
table highs beneath uplands to hydrologic sinks in the lowlands (modified from Hubbert 1940); (b) local flow systems, which
are recharged at water table highs and discharge to adjacent lowlands, overlie a regional flow system (modified from Winter
1976); (c) ground water flow systems where water table highs do not underlie topographic highs. The surface water bodies are
flow-through with respect to ground water.

990 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000


989-1000 GW Watershed 03 12/1/03 4:53 PM Page 991

recharge and discharge conditions. The superposition of distant source, where the ground water passed beneath one
flow systems can result in ground water discharge to a sur- or more local flow systems before discharging to the sur-
face water body from more than one flow system; deeper face water.
flow systems may have their watershed far beyond the sur- A number of case studies are presented in this paper
face watershed (Tóth 1962; Haitjema 1995). For example, that demonstrate the difficulties in determining the extent
a stream located at the lowest point (left side) of the hydro- of ground water watersheds in various types of terrain that
logic sections shown in Figures 1b and 1c would receive represent various combinations of high and low permeabil-
ground water inflow from the local flow system underlying ity and low and high regional topographic relief. Determin-
the contiguous surface watershed, but it also would receive ing the extent of ground water watersheds by assuming the
ground water from the regional flow system that extends configuration of the water table reflects the configuration of
the entire length of the section. If the ground water dis- land surface is particularly difficult in hummocky terrain
charge from the regional flow system is substantial, of having relatively low regional topographic relief, such as
undesirable quality, or contaminated, considering ground glacial and dune terrain outside of mountainous areas. Even
water only from the local flow system contiguous to the surface watersheds can be difficult to define in these types
stream could lead to ineffective management because the of terrain because the topographic highs have highly irreg-
recharge area of the regional flow system extends beyond ular shapes, they can be very subtle, and there can be many
the local surface watershed. topographically closed depressions that do not contribute
Surface watersheds have numerous variations in size, surface runoff to the larger drainage basin. Closed depres-
shape, relief, geologic substrate, and rates of water runoff. sions in hummocky terrain commonly contain lakes and
Similarly, ground water flow systems also have numerous wetlands that are in contact with, and therefore are part of,
variations in size, shape, geology, and rates of ground water ground water flow systems (Winter 1999).
movement. It is beyond the scope of this paper to try to
describe all of the possible variations in comparing surface Terrain with High Permeability
and ground water watersheds; however, two issues that and Low Regional Topographic Relief
watershed researchers need to be aware of when dealing Terrain that is relatively permeable, has low regional
with sources of ground water in watersheds are examined topographic slope, and is locally hummocky may have
in this paper. The issues are (1) the difficulties in determin- extensive ground water flow systems. The flow systems
ing the extent of ground water watersheds, and (2) the begin near regional water table highs and terminate at
dynamic characteristics of ground water divides in some regional lows. Lakes and wetlands that lie within these flow
hydrogeologic settings. This paper focuses mostly on stud- systems can be thought of as ground water outcrops, where
ies where a considerable amount of effort went into deter- ground water moves into the lake or wetland on the upgra-
mining the ground water component of the hydrologic sys- dient side and lake or wetland water recharges ground
tem of small watersheds. In addition, the focus is primarily water on the downgradient side. Thus, the contributing area
on unconfined ground water flow systems in unconsoli- of ground water to a lake or wetland selected for a research
dated deposits and/or shallow bedrock associated with such watershed can range widely, depending on where it is
deposits. located within the ground water flow system. The con-
That ground water flow divides do not coincide with tributing area for ground water to the watershed would be
surface watersheds in areas of confined aquifers is well relatively small if it is located near the upper end of the
known, and has been documented for a wide range of flow system and relatively large if it is located near the
scales. For example, Winograd and Thordarson (1975) lower end. However, except for watersheds on ground
described a large-scale interbasin regional flow system as water divides, the contributing area of ground water to the
the source of water to Ash Meadows in Nevada, and Hunt lake or wetland is almost certain to extend beyond the local
et al. (2001) determined that the ground water basin of a surface watershed. The following case studies of terrain
spring in Wisconsin did not coincide with the surface having high permeability and low regional topographic
watershed of the spring. Because ground water flow in relief are taken from field sites in glacial terrain in Wiscon-
many confined aquifers is unrelated to surface watersheds, sin and Minnesota, and dune terrain in Nebraska.
such settings are not discussed in this paper.
Trout Lake Area in Wisconsin
The Trout Lake area lies within the north-central high-
Determining the Extent lands in Wisconsin (Figure 2). Glacial deposits that are ~40
of Ground Water Watersheds to 60 m thick and that consist largely of outwash sand and
Two characteristics of ground water flow systems gravel underlie the area. The area has many topographically
make it difficult to determine the extent of ground water closed depressions that contain lakes or wetlands and that
watersheds. (1) The configuration of the water table com- have small surface watersheds. To determine the relation
monly does not reflect the configuration of land surface; between ground water and the lakes that lie within the
therefore, the boundaries of a shallow ground water flow Trout Lake watershed, Cheng (1994) developed a steady-
system defined on the basis of surface topography may not state, three-dimensional, finite-difference model. The
reflect the true boundaries of the flow system. (2) Ground model boundaries were defined on the basis of surface
water flow systems of different magnitude can overlie one topography. Because analytic element models have proved
another, as shown in Figures 1b and 1c; therefore, ground useful in solving for the boundary conditions of ground
water discharging to a surface water body may be from a water flow systems (Haitjema 1995; Kelson 1998), Hunt et
T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000 991
989-1000 GW Watershed 03 12/1/03 4:53 PM Page 992

from a large east-west-trending moraine in north-central


Minnesota (Figure 2). The area is underlain by glacial
deposits that are >120 m thick, and consist of thick alter-
nating units of till and sand and gravel (Winter and Rosen-
berry 1997). The surficial deposits are fluvial silt, sand, and
gravel. The area has many topographically closed depres-
sions that have small surface watersheds (Figure 4a).
Williams Lake was initially chosen for detailed study of
water and chemical budgets because it is a topographically
closed basin. However, installation of wells between Crys-
tal, Williams, and Mary lakes indicated that a ground water
divide did not underlie the surface watershed, and that
Williams Lake receives water that seeps out of Crystal
Lake and loses water that eventually seeps into Mary Lake.
Thus, the lake lies within a ground water flow system that
extends from Crystal Lake to Mary Lake. The exact loca-
tion of the southern ground water divide in the Crystal Lake
area has never been determined; i.e., it is not known if a
ground water divide is present beneath the topographic
ridge on the south side of Crystal Lake or if the lake itself
forms part of the divide.
Figure 2. Location of watersheds in the United States, Early in the study of the Williams Lake area, it was
Canada, Costa Rica, and Germany discussed in this paper. discovered that ground water flow systems of different
magnitude were superimposed. A numerical cross-sec-
tional model of an area extending from the ground water
divide east of Williams Lake to Mary Lake indicated that
al. (1998) re-evaluated the boundary conditions of the only the shallower ground water in the surficial aquifer dis-
ground water flow system that was used for the finite-dif- charged to Williams Lake. Deeper ground water within the
ference model of the Trout Lake area, using an analytic ele- surficial aquifer passed at depth beneath the lake and dis-
ment model. Comparison of the boundary based on surface charged to Mary Lake (Siegel and Winter 1980).
topography and that based on the analytic element model is Evidence that ground water divides do not underlie
shown in Figure 3. The ground water basin defined by the surface divides, and that flow systems of different magni-
analytic element model is larger and has a different shape, tude overlie one another, is present elsewhere in the
compared to the finite-difference model. The improved def- Shingobee River headwaters area. For example, the large
inition of boundary conditions resulted in improvement of
overall ground water flux calculations and distribution of
fluxes in the ground water basin.
The Trout Lake area also provides an example of how
considering the superposition of flow systems can lead to
increased understanding of ground water relations to sur-
face water bodies in small watersheds. One of the initial
studies of lake and ground water interaction in the Trout
Lake area focused on ground water flow and geochemical
processes between Crystal and Big Muskellunge lakes
(Kenoyer and Anderson 1989; Kim et al. 1999). In a later
study, which involved sampling ground water from deeper
wells in the vicinity, chemical data indicated that a deeper
ground water flow system from upgradient of Crystal Lake
underlay the shallower ground water moving from Crystal
Lake to Big Muskellunge Lake (Hunt et al., in review). A
more regional ground water model of the entire area
between Crystal Lake and the Allequash River was needed
to simulate the water age and source that was measured at
depth (Pint et al. 2003).

Shingobee River Headwaters Area in Minnesota


Figure 3. Trout Lake area in Wisconsin showing the differ-
The Shingobee River headwaters area in Minnesota ence in a ground water divide selected from surface topogra-
provides another example of the difficulties in defining phy (assuming ground water divides underlie topographic
ground water divides associated with small watersheds in divides) and a ground water divide determined by an analytic
element (AE) screening model. (Modified from Hunt, Ander-
glacial terrain. The Shingobee River headwaters area is
son, and Kelson 1998)
located on a small topographic ridge that extends south
992 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000
989-1000 GW Watershed 03 12/1/03 4:53 PM Page 993

m/year), annual recharge within the local watershed of


Shingobee Lake is ~280,000 m3/year; whereas the excess
of surface outflow over surface inflow to the lake is more
than 1 million m3/year.
In an attempt to obtain an overall picture of ground
water and surface water interaction within the entire Shin-
gobee River headwaters area, a three-dimensional ground
water flow model was developed for the surficial aquifer
(Filby et al. 2002). The model results indicated that some of
the intermediate-scale flow systems that discharge into Shin-
gobee Lake and Little Shingobee Fen may be recharged as
far away as the Williams and Crystal lakes area.

Island Lake Area in Nebraska


Lakes and wetlands occupy many of the interdunal
lowlands in the sandhills of Nebraska (Figure 2). Similar to
the studies in the Shingobee River headwaters area in Min-
nesota, construction of numerous wells between Island
Lake and neighboring lakes indicated that ground water
divides do not commonly underlie surface watershed
divides. The configuration of the water table in the vicinity
of this group of lakes indicates that ground water moves
generally from north to south and that the lakes are in direct
connection with the ground water system (Figure 5) (Win-
ter 1986). The group of lakes studied in this part of the
Crescent Lake National Wildlife Refuge lie near the lowest
part of a vast ground water flow system that may have its
regional divide many kilometers to the north. Although
water moves from lake to lake through the ground water
Figure 4. Shingobee River headwaters area in Minnesota: (a)
map showing surface water bodies, surface watersheds, trace system, chemical data indicate that the water does not move
of hydrologic section W-E, and direction of ground water as an unmodified lake-water plume from lake to lake
flow between Crystal, Williams, and Mary lakes in the upper (LaBaugh 1986). Near-shore seepage of ground water into
(southern) part of the area; (b) hydrologic section W-E show- these lakes on their upgradient side does not reflect the
ing local and regional ground water flow systems in the Shin- chemistry of the next lake upgradient; instead, it has chem-
gobee River, Little Shingobee Fen, and Little Shingobee Lake
area in the lower (northern) part of the area. ical characteristics of freshly recharged ground water.
Therefore, ground water inflow to a given lake is a combi-
nation of ground water that was recharged in the uplands
between the lakes and by upgradient lake water that moves
surface discharge from the Little Shingobee Fen (~360,000 deeper through the ground water system. In addition, it is
m3/year) can be accounted for only by having ground water likely that deeper ground water may bypass some of the
inflow to the fen be derived from an intermediate-scale lakes at depth beneath the shallower ground water flow sys-
ground water flow system. (If all of the recharge that took tems, as at the Trout Lake (Wisconsin) and Williams Lake
place in the local watershed of the fen [~0.13 m/year] dis- (Minnesota) field sites.
charged to the fen, it would amount to only ~12,000
m3/year.) The intermediate flow system is recharged 1 to 2 Terrain Having Low Permeability
km east of the Shingobee River and passes beneath the river and Low Regional Topographic Relief
as well as the local flow system associated with the Shin- The configuration of the water table in terrain having
gobee River (Figure 4b) (Winter et al. 2001). low permeability, such as glacial till, generally is more
Shingobee Lake must also be receiving ground water complex than in the more permeable sand-and-gravel set-
inflow from intermediate and regional ground water flow tings presented in the previous section of this paper. In
systems. The surface watershed of Shingobee Lake is rela- some areas of fine-grained till, water table mounds can be
tively small, yet numerous springs are present around its present beneath topographic highs on a local scale, result-
perimeter and in the lakebed. Some of the springs have very ing in some lakes and wetlands being discharge areas for
stable flow relatively independent of climate variability, local ground water flow systems. Alternatively, the depres-
which indicates that their source is a relatively large and sions themselves can be the focal points for ground water
stable ground water flow system. In addition, ground water recharge, where the water table mounds underlie the
supplies about a third of the inflow to Shingobee Lake depressions rather than the topographic highs (Lissey
(Rosenberry et al. 1997), which is much more water than 1971). As a result of the complex configuration of the water
could be supplied from ground water recharged within its table, individual surface water bodies in low-permeability
surface watershed. Assuming the same amount of recharge terrain can be recharge areas, discharge areas for closed
as was used for the Little Shingobee Fen calculation (~ 0.13 local flow systems, have flow-through conditions similar to
T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000 993
989-1000 GW Watershed 03 12/1/03 4:54 PM Page 994

ground water from beyond their local watersheds and have


seepage to ground water. Furthermore, superposition of
ground water flow systems also is common here, and
throughout the Missouri Coteau (Swanson et al. 1988).
However, unlike the sites discussed in the previous section,
the low permeability of the till results in much more relief
of the water table, which in turn results in the presence of
ground water divides beneath some of the local surface
watershed divides.

Harp Lake Catchment in Ontario, Canada


Harp 4–21 is a small (3.7 ha) headwater catchment
within the Harp Lake watershed in southern Ontario. The
catchment is near the southern margin of the Canadian
Shield (Figure 2), and the surface deposit overlying the
crystalline bedrock is glacial till that ranges in thickness
from 1 to 13 m (Hinton et al. 1993). Topographic relief of
the catchment is ~38 m. Detailed study of ground water
movement in this small catchment indicated that the water
table was not a subdued replica of the land surface and that
water table contours were not parallel to land surface con-
tours. The study indicated that the surface watershed area
differed by as much as 57% from the ground water water-
shed area of the stream.

Terrain Having Low Permeability


and High Topographic Relief
Figure 5. Island Lake area in Nebraska showing surface The studies discussed previously provide examples of
water bodies, surface watersheds, contours of the water table the difficulties in defining the boundaries of ground water
(m), and direction of ground water flow on October 28, 1982. watersheds associated with surface watersheds in relatively
(Modified from Winter 1986) flat landscapes. However, even in areas of high relief (hun-
dreds of meters) that have well-defined surface watersheds,
ground water divides may differ from surface watersheds.

lakes and wetlands in permeable terrain, or be discharge


areas for local, intermediate, and regional ground water
flow systems. The following case studies of terrain having
low permeability and low regional topographic relief are
taken from field sites in glacial terrain in North Dakota and
Ontario, Canada. Hydraulic conductivity of the glacial till
at both sites is highly variable, but at both sites it averages
about 10–7 m/sec.

Cottonwood Lake Area in North Dakota


The Cottonwood Lake area is situated on one of the
highest parts of the eastern edge of a large moraine, the
Missouri Coteau, in eastern North Dakota (Figure 2). The
glacial deposits in the Cottonwood Lake area are as much
as 140 m thick and consist predominantly of clayey, silty
till. The area has many steep-sided closed depressions,
most of which contain wetlands that have small surface
watersheds (Figure 6). Although Wetland P1 was the pri-
mary focus of study with respect to determining detailed
water and chemical budgets, installation of wells through-
out the Cottonwood Lake area indicated that the wetland
receives ground water from beyond its local surface water-
shed. The ground water flow system extends to the topo- Figure 6. Map of the Cottonwood Lake area in North Dakota
graphic high in the southeastern part of the area, which is showing wetlands, surface watersheds, configuration of the
underlain by a regional water table high. The Cottonwood water table, and direction of ground water flow in the vicin-
Lake area is similar to the sites discussed in the previous ity of Wetland P1 on May 31, 1987. (Modified from Winter
and Rosenberry [1995]).
section of this paper in that many of the wetlands receive

994 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000


989-1000 GW Watershed 03 12/1/03 4:54 PM Page 995

The following case studies of terrain having low perme-


ability and high relief are taken from field sites in moun-
tainous terrain in New Hampshire and Costa Rica.

Mirror Lake Area in New Hampshire


The Mirror Lake watershed is located near the lower
end of the Hubbard Brook valley in the White Mountains of
central New Hampshire (Figure 2). The lake, which lies at
an altitude of 213 m, is bounded to the west and north by
steep hills that reach a maximum altitude of 469 m (Figure
7). The watershed is underlain by fractured crystalline
bedrock, which is overlain by glacial deposits. Thickness of
the glacial deposits varies from zero to a few meters in the
higher part of the Mirror Lake watershed to as much as 30
m in the lower part. The glacial deposits throughout most
of the Mirror Lake watershed consist primarily of silty,
sandy till, containing numerous cobbles and boulders.
A preliminary cross-sectional numerical model of the
Mirror Lake area (Harte and Winter 1996) indicated that
the ground water divide of the flow system associated with
Mirror Lake was located beyond the surface watershed as
well as beyond the watershed of Norris Brook, the next
stream up the valley. The model results indicated that part
of the ground water flow system associated with Mirror
Lake passes beneath a small local flow system associated
with Norris Brook. The model results also indicated that a
deeper ground water flow system passes beneath the Mirror
Lake ground water system and discharges to the Pemige- Figure 7. Mirror Lake area in New Hampshire showing the
wassett River. surface watershed, and the ground water basin associated
To provide a more complete picture of Mirror Lake’s with the lake as determined by a three-dimensional ground
ground water basin, a three-dimensional numerical model water flow model. (Modified from Tiedeman et al. 1997)
was constructed by Tiedeman et al. (1997). The model
results also indicated that the Mirror Lake ground water
basin extends beyond Norris Brook. By mapping the three- tributed as much as 84% of the water that discharged as
dimensional flow system, the model was able to determine riparian seeps. Because of the difference in solutes between
the areal extent of the ground water basin that is recharged the shallow local ground water and the deeper regional
at the water table, and the areal extent of the part that moves ground water, the contribution of solutes by interbasin
as underflow beneath Norris Brook (Figure 7). The model transfer was even higher than it was for the interbasin trans-
also determined that the ground water basin is 50% larger fer of water. For example, at the location and time that 49%
than the surface watershed. of the stream water was ground water derived from distant
sources, the chloride contributed by those sources was
La Selva Biological Station in Costa Rica 92%. In the case of the seeps, it was found that as much as
The La Selva Biological Station lies in the transition 99% of the chloride was brought into the area by ground
zone between the steep foothills of the Cordillera Central water from distant sources. By correlation of chloride with
and the Caribbean coastal plain in Costa Rica (Figure 2). other major ions, the data indicated that most solutes were
Bedrock in the area is largely volcanic in origin. The bio- brought into the area by ground water from distant sources.
logical station is operated by the Organization for Tropical A major finding of the study was that the ground water
Studies and includes a number of small watersheds that are derived from outside the basin had a substantial effect on
used for research and educational purposes. Using chemi- the chemistry of not only the streams but also much of the
cal data from more than 800 samples of surface water and shallow ground water near streams (Figure 8).
ground water, Genereux et al. (2002) found that surface
water and shallow ground water contained two chemically Dynamic Characteristics of Ground Water Divides
distinct waters. One water type had low solutes and is typ- Unconfined ground water is recharged by infiltration
ical of the water that drains from hillslope soils within the of precipitation and movement of water through the unsat-
surface watersheds following rainfall. The second water urated zone to the water table. The areal distribution of
type was more mineralized and was believed to be ground recharge can be highly variable depending on the distribu-
water derived from beyond the local surface watersheds. tion of precipitation, permeability of geologic deposits, and
Chemical mixing calculations indicated that as much configuration of the land surface. Permeability of the geo-
as 49% of the water in some streams during baseflow con- logic deposits controls the rates of infiltration and distribu-
ditions consisted of deep ground water that had moved into tion of water in the subsurface. Configuration of the land
the area from distant sources. This source of water also con- surface affects the areal distribution of recharge because,

T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000 995


989-1000 GW Watershed 03 12/1/03 4:54 PM Page 996

for a given permeability, water moving downward through the ground water watershed of Crane Lake included Island
the unsaturated zone reaches the water table sooner where Lake. During the high-water period of 1983, a water table
the unsaturated zone is thinner. Hence, recharge commonly trough was located in the vicinity of well 8 (Figure 9f).
is focused initially beneath depressions in the land surface
in some landscapes (Lissey 1971) and near surface water Cottonwood Lake Area in North Dakota
bodies that are hydraulically connected to ground water The general hydrogeology of the Cotttonwood Lake
(Winter 1983). Because precipitation is the ultimate source area was described in a previous section of this paper. The
of recharge, the temporal and spatial configuration of the ground water divide between local flow systems associated
water table for a given landform and geology depends on with individual wetlands of interest for this paper lies
the temporal and spatial distribution of precipitation. The between Wetlands T1 and P8 (Figure 6). At this locality, a
result is that the water table and ground water divides move ground water divide underlies the surface divide most of the
up and down, and divides may move laterally as well. time. However, during very dry conditions, the water table
Three examples of field studies are presented where the mound on the west side of Wetland T1 occasionally lowers
movement of ground water divides has been documented. enough for ground water to move from the ground water
The sites are in Nebraska, North Dakota, and Germany. watershed of Wetland P1 to the ground water watershed of
Wetland P8 (Winter and Rosenberry 1995). During those
Island Lake Area in Nebraska times, the ground water watershed for the local flow system
The general hydrogeology of the Island Lake area was contributing ground water to Wetland P1 becomes part of
described in the previous section of this paper. One aspect the ground water watershed of Wetland P8, greatly increas-
of the studies at Island Lake was to examine the recharge ing the extent of the local ground water flow system con-
process beneath the sand dunes. Water level data from lines tributing to Wetland P8. Since 1980, this condition occurred
of observation wells installed between the lakes indicated during a one-year drought in 1985 and the five-year drought
that the divides, both water table mounds and water table from 1988 to 1992.
troughs, moved laterally depending on wet or dry condi-
tions (Winter 1986). Where water table mounds were pre- Lake Stechlin Area in Germany
sent beneath a hummocky dune between Hackberry and Lake Stechlin lies in the lake district of northeast Ger-
Island lakes, the divide moved from the vicinity of well 13 many (Figure 2). It and nearby Lakes Dagow, Peetsch, and
during the high-water period of 1982 (Figure 9a) to well 2 Glietzen receive ground water inflow but have no perennial
during the low-water period of 1982 (Figure 9b), and back stream inflow. To gain an understanding of the extent of the
to well 13 during the high-water period of 1983 (Figure 9c). ground water basins that contribute water to these lakes,
In a later study, Keen (1992) documented the dynamic Holzbecher (2001) developed a two-dimensional numerical
characteristics of the water table divide between Island and ground water model of the area, focusing particularly on
Hackberry lakes by installing additional water table wells the ground water systems between the four lakes. The
and instrumenting them with continuous recorders. model was calibrated to lake levels, which have been mea-
On the opposite side of Island Lake, where the line of sured since 1901, and ground water levels, which have been
observation wells traversed a sharp-crested dune, the move- measured at ~20 wells since 1959. Results of simulations
ment of a water table trough also was dynamic. During the for wet, dry, and very dry conditions are shown in Figure
high-water period of 1982, the nadir of the trough was in 10.
the vicinity of well 7 (Figure 9d). During the low-water For wet conditions (Figure 10A), a ground water
period of 1982, there was no water table trough between divide is present between lakes Stechlin and Glietzen and
Island and Crane lakes (Figure 9e). Under this condition, between lakes Stechlin and Peetsch. A divide is not present
between lakes Dagow and Stechlin, permitting seepage of
water from Lake Dagow to move as ground water directly
to Lake Stechlin.
For dry conditions (Figure 10b), the divide between
lakes Stechlin and Glietzen moved closer to Lake Stechlin
compared to the wet conditions; however, it is breached in
a very small area, permitting a small amount of seepage
from Lake Stechlin to move as ground water directly to
Lake Glietzen. The divide between lakes Stechlin and
Peetsch moved closer to Lake Stechlin compared to the wet
conditions. The dimensions of the ground water flow sys-
tem between lakes Dagow and Stechlin changed little com-
pared to the wet conditions.
For very dry conditions (Figure 10c), the divide
between lakes Stechlin and Glietzen moved even closer to
Lake Stechlin compared to the dry conditions. Further-
Figure 8. Hydrologic section in the La Selva area in Costa more, the breach in the ground water divide became larger,
Rica showing the ratio of local ground water to ground water
that had moved into the area from outside the watershed
permitting a larger amount of seepage from Lake Stechlin
(fwater). (Modified from Genereux et al. 2002) to move as ground water directly to Lake Glietzen. The
divide between lakes Stechlin and Peetsch moved closer to
996 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000
989-1000 GW Watershed 03 12/1/03 4:54 PM Page 997

Figure 9. Hydrologic sections in the Island Lake area in Nebraska showing water table profiles between Hackberry and Island
lakes (a, b, and c), and between Island and Crane lakes (d, e, f) for three dates. (Modified from Winter 1986)

Lake Stechlin compared to the dry conditions. The dimen- and are surface expressions of ground water flow systems
sions of the ground water flow system between lakes that may be much larger than the surface watersheds. If a
Dagow and Stechlin changed considerably compared to the lake or wetland is located near the upper end of an extensive
dry conditions. In this area, the much larger ground water ground water flow system, it will have a smaller ground
flow system that moved water from Lake Dagow to Lake water watershed than if it is located near the lower end of the
Stechlin greatly reduced the area of the flow system that flow system. Chemical budgets of topographically isolated
moved water to Lake Stechlin from the area between lakes surface water are related to the amount of water and solutes
Stechlin and Peetsch. received from ground water and the amount of water and
solutes lost to ground water (Schwartz and Gallup 1978;
Wood and Sanford 1990; Gosselin et al. 1994; Cheng and
Discussion Anderson 1994; LaBaugh et al. 2000). Therefore, a surface
Some general observations can be derived from the water body near the upper end of a flow system would be
results of the studies presented in this paper. For example, if expected to have fewer solutes contributed by ground water
a region has a relatively permeable unconfined aquifer, (1) than a surface water body near the lower end. As pointed out
water table divides (boundaries of local ground water flow) by Winter (2001), this is the case for the Trout Lake area in
commonly do not underlie local surface watersheds, and (2) Wisconsin and the Shingobee River headwaters area in
regional ground water flow systems can underlie local flow Minnesota. However, it is not the case for the Island Lake
systems. Examples of such types of terrain include outwash area in Nebraska. Here, the lakes are increasingly less min-
sand and gravel, dune sand, beach sand, and fluvial sand and eralized, as they are positioned lower in the ground water
gravel formations. Even though surface water bodies such flow system. This seeming paradox brings out the impor-
as lakes and wetlands may lie within closed surface water- tance of considering the relative contributions of superim-
sheds, they generally are in direct contact with ground water posed flow systems.

T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000 997


989-1000 GW Watershed 03 12/1/03 4:54 PM Page 998

body accounts for the greater mineralization of water bod-


ies lower in the ground water flow system. In the case of the
Nebraska site, it is hypothesized that locally recharged
fresh ground water between the lakes contributes more
water to the lakes than the more mineralized deeper ground
water. The deeper ground water eventually discharges to
major regional hydrologic drains—a large wetland that
forms the headwaters of the Blue River, a tributary to the
North Platte River, and to the North Platte River itself.
If a region is underlain by poorly permeable geologic
deposits, such as clayey till and lacustrine and fluvial silt
and clay, ground water flow system divides commonly
underlie local surface watersheds. However, the ground
water divides may not completely enclose the surface water
body. At the gaps in the divide, ground water from beyond
the surface watersheds can discharge to the surface water
body. Even though the movement of the shallowest ground
water from surface water body to surface water body down
a regional ground water gradient is less common and more
disjointed in areas of poorly permeable deposits, superim-
posed flow systems have an important role in determining
the quantity and types of solutes in the surface water. In the
case of the North Dakota site, the deeper flow systems result
in the surface water bodies in lower parts of the region being
more mineralized than those in the higher parts.
The dynamics of ground water flow system divides
pose a challenging dilemma to research in small water-
sheds. Few studies have examined the effect of moving
divides on the fluxes of solutes within the ground water
system or on their discharge to surface water. One can only
speculate on the effect of moving ground water divides on
fluxes to surface water, based on the three studies presented
in the previous section of this paper. The shifting divides
between Island Lake and its two neighbors, Hackberry and
Crane lakes, probably would not affect the solute concen-
trations of the lakes because the distance of the shifts are
not great and the divides remain in the same general vicin-
ity. Furthermore, the lines of section shown are aligned
roughly perpendicular to the regional flow system; there-
fore, the gradients roughly perpendicular to the sections are
greater than those along the sections. The vast majority of
seepage to and from the lakes is on their upgradient and
downgradient sides.
In the case of the North Dakota site, the ground water
Figure 10. Maps of ground water divides and direction of divide between Wetlands P1 and P8 disappears only during
ground water flow in the vicinity of Lake Stechlin and nearby the most severe droughts, and has a recurrence interval of
lakes in northeast Germany for (a) wet, (b) dry, and (c) very years to decades. Furthermore, even during those times
dry conditions. (Modified from Holzbecker 2001) when the divide is not present, the rates of ground water
movement are so low in the poorly permeable till that large
masses of solutes probably do not move from the ground
Most of the case studies presented in this paper indi- water watershed of Wetland P1 to that of Wetland P8.
cate that ground water flow systems of different magnitude In the case of the Lake Stechlin area, the moving
are superimposed in the general study area. The sites under- ground water divides could be fairly important to the water
lain by permeable geologic deposits generally have a slop- and solute fluxes to the lakes. The divides move a relatively
ing water table connecting the lakes and wetlands, indicat- large distance, particularly between Lake Stechlin and
ing that the shallowest ground water moves from surface Lakes Dagow and Peetsch.
water body to surface water body. These sites also have
deeper ground water flow systems that move water from
higher parts of the flow system to lower parts, sometimes Conclusion
bypassing some of the surface water bodies. In many cases, Researchers studying small watersheds need to be
discharge of this deeper ground water to a surface water aware that ground water flow divides do not underlie
998 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000
989-1000 GW Watershed 03 12/1/03 4:54 PM Page 999

surface divides in many settings. Only if a surface water- groundwater flowpaths and the need for multiple criteria. In
shed of a research site is at the highest ridge away from review.
major hydrologic sinks such as regional rivers, can one be Hunt, R.J., J.J. Steuer, M.T.C. Mansor, and T.D. Bullen. 2001.
Delineating a recharge area for a spring using numerical
sure that ground water is not moving into the area from dis- modeling, Monte Carlo techniques, and geophysical investi-
tant sources. For most other watersheds, ground water can gation. Ground Water 39, no. 5: 702–712.
move into watersheds either because ground water divides Hurley, J.P., D.E. Armstrong, G.J. Kenoyer, and C.J. Bowser.
are not present beneath the surface divides or because 1985. Ground water as a silica source for diatom production
ground water moves into the area from deeper flow sys- in a precipitation-dominated lake. Science 227, 1576–1578.
Hynes, H.B.N. 1975. The stream and its valley. Verhandlungen
tems. Ground water divides move in response to changing Internationale Vereinigung Limnologie 19, 1–15.
recharge conditions, which in turn is related to the dynam- Keen, K.L. 1992. Geomorphology, recharge, and water-table
ics of climate and precipitation. fluctuations in stabilized sand dunes: The Nebraska sand
hills, U.S.A. Ph.D. diss., Department of Geology and Geo-
physics, University of Minnesota.
Kelson, V.A. 1998. Practical advances in groundwater modeling
References with analytic elements. Ph.D. diss., School of Public and
Baedecker, M.J., and L.C. Friedman. 2000. Water, energy, and Environmental Affairs, University of Indiana.
biogeochemical budgets: A watershed research program. Kenoyer, G.J., and M.P. Anderson. 1989. Groundwater’s
USGS Fact Sheet 165–99. dynamic role in regulating acidity and chemistry in a pre-
Cheng, X. 1994. Numerical analysis of groundwater and lake sys- cipitation-dominated lake. Journal of Hydrology 109,
tems with application to the Trout River basin, Vilas County, 287–306.
Wisconsin. Ph.D. diss., Department of Geology and Geo- Kim, K., M.P. Anderson, and C.J. Bowser. 1999. Model calibra-
physics, University of Wisconsin-Madison. tion with multiple targets: A case study. Ground Water 37,
Cheng, X., and M.P. Anderson. 1994. Simulating the influence of no. 3: 345–351.
lake position on groundwater. Water Resources Research 30, LaBaugh, J.W. 1986. Limnological characteristics of selected
no. 7: 2041–2049. lakes in the Nebraska sandhills, U.S.A., and their relation to
Filby, S.K., S.M. Locke, M.A. Person, T.C. Winter, D.O. Rosen- chemical characteristics of adjacent ground water. Journal
berry, J.L. Nieber, W.J. Gutowski, and E. Ito. 2002. Mid- of Hydrology 86, 279–298.
Holocene hydrologic model of the Shingobee watershed, LaBaugh, J.W., D.O. Rosenberry, and T.C. Winter. 1995.
Minnesota. Quaternary Research 58, no. 3: 246–254. Groundwater contribution to the water and chemical bud-
Genereux, D.P., S.J. Wood, and C.M. Pringle. 2002. Chemical gets of Williams Lake, Minnesota, 1980–1991. Canadian
tracing of interbasin groundwater transfer in the lowland rain- Journal of Fisheries and Aquatic Sciences 52, 754–767.
forest of Costa Rica. Journal of Hydrology 258, 163–178. LaBaugh, J.W., T.C. Winter, and D.O. Rosenberry. 2000. Com-
Goodrich, D.C., and D.A. Woolhiser. 1993. Catchment hydrol- parison of the variability in fluxes of ground water and
ogy. U.S. National Report to International Union of Geo- solutes in lakes and wetlands in central North America. Ver-
desy and Geophysics. Reviews of Geophysics Supplement, handlungen Internationale Vereinigung Limnologie 27,
202–209. 420–426.
Gosselin, D.C., S. Sibray, and J. Ayers. 1994. Geochemistry of Leopold, L.B. 1974. Water–A Primer. San Francisco: W.H.
K-rich alkaline lakes, western Sandhills, Nebraska. Freeman and Co.
Geochimica et Cosmochemica Acta 58, 1403–1418. Likens, G.E. 1984. Beyond the shoreline: A watershed-ecosys-
Haitjema, H.M. 1995. Analytic Element Modeling of Groundwa- tem approach. Verhandlungen Internationale Vereinigung
ter Flow. San Diego, California: Academic Press. Limnologie 22, 1–22.
Harte, P. T., and T.C. Winter. 1996. Factors affecting recharge to Likens, G.E., F.H. Bormann, N.M. Johnson, and R.S. Pierce.
crystalline rock in the Mirror Lake area, Grafton County, 1967. The calcium, magnesium, potassium, and sodium
New Hampshire. In USGS Toxic Substances Hydrology budgets for a small forested ecosystem. Ecology 48,
Program—Proceedings of the Technical Meeting, Colorado 772–785.
Springs, Colorado, September 20–24, 1993, ed. D.W. Mor- Lissey, A. 1971. Depression-focused transient ground water flow
ganwalp and D.A. Aronson, 141–150. USGS Water- patterns in Manitoba. Geological Association of Canada
Resources Investigations Report 94–4015. Special Paper 9, 333–341.
Hemond, H.F. 1980. Biogeochemistry of Thoreau’s Bog, Con- McCammon, B.J., J. Rector, and K. Gebhart. 1998. A framework
cord, Massachusetts. Ecological Monographs 50, 507–526. of analyzing the hydrologic condition of watersheds. U.S.
Hewlett, J.D., H.W. Lull, and K.G. Reinhart. 1969. In defense of Bureau of Land Management Technical Notes 405.
small watersheds. Water Resources Research 5, 306–316. Moore, W.S. 1999. The subterranean estuary: A reaction zone of
Hinton, M.J., S.L. Schiff, and M.C. English. 1993. Physical prop- ground water and sea water. Marine Chemistry 65, 111–
erties governing groundwater flow in a glacial till catch- 125.
ment. Journal of Hydrology 142, 229–249. Pint, C.D., R.J. Hunt, and M.P. Anderson. 2003. Flow path delin-
Holmes, R.M. 2000. The importance of ground water to stream eation and ground water age, Allequash Basin, Wisconsin.
ecosystem function. In Streams and Ground Waters, ed. Ground Water, this issue.
R.M. Holmes and P.J. Mulholland, 137–148. San Diego, Rosenberry, D.O., T.C. Winter, D.A. Merk, G.H. Leavesley, and
California: Academic Press. L.D. Beaver. 1997. Hydrology of the Shingobee River head-
Holzbecker, E. 2001. The dynamics of subsurface water divides: waters area. In Hydrological and Biogeochemical Research
Watersheds of Lake Stechlin and neighboring lakes. Hydro- in the Shingobee River Headwaters Area, North-Central
logical Processes 15, 2297–2304. Minnesota, ed. T.C. Winter, 19–23. USGS Water-Resources
Hubbert, M.K. 1940. The theory of ground water motion. Jour- Investigations Report 96–4215.
nal of Geology 48, 785–944. Schindler, D.W., R.W. Newberry, K.G. Beaty, and P. Campbell.
Hunt, R.J., M.P. Anderson, and V.A. Kelson. 1998. Improving a 1976. Natural water and chemical budgets for a small Pre-
complex finite-difference ground water flow model through cambrian lake basin in central Canada. Journal of the Fish-
the use of an analytic element screening model. Ground eries Research Board of Canada 33, 2526–2543.
Water 36, no. 6: 1011–1017. Schwartz, F.W., and D.N. Gallup. 1978. Some factors controlling
Hunt, R.J., D.P. Krabbenhoft, T.D. Bullen, J.F. Walker, C. the major ion chemistry of small lakes: Examples from the
Kendall, and M.P. Anderson. Difficulties in determining prairie parkland of Canada. Hydrobiologia 58, 65–81.

T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000 999


989-1000 GW Watershed 03 12/1/03 4:54 PM Page 1000

Siegel, D.I., and T.C. Winter. 1980. Hydrologic setting of Winter, T.C. 1986. Effect of ground water recharge on configu-
Williams Lake, Hubbard County, Minnesota. USGS Open- ration of the water table beneath sand dunes and on seepage
File Report 80–403. in lakes in the sandhills of Nebraska. Journal of Hydrology
Sosin, A., D. Brady, M. McCarthy, W. Cooter, and S. Alexander. 86, 221–237.
1995. Watershed protection: A project focus. U.S. Environ- Winter, T.C. (ed.). 1997. Hydrological and biogeochemical
mental Protection Agency Report EPA841-R–95–003. research in the Shingobee River headwaters area, north-cen-
Swanson, G.A., T.C. Winter, V.A. Adomaitis, and J.W. tral Minnesota. USGS Water-Resources Investigations
LaBaugh. 1988. Chemical characteristics of prairie lakes in Report 96–4215.
south-central North Dakota: Their potential for influencing Winter, T.C. 1999. Relation of streams, lakes, and wetlands to
use by fish and wildlife. U.S. Fish and Wildlife Technical groundwater flow systems. Hydrogeology Journal 7, 28–45.
Report 18. Winter, T.C. 2001. The concept of hydrologic landscapes. Jour-
Tiedeman, C.R., D.J. Goode, and P.A. Hsieh. 1997. Numerical nal of the American Water Resources Association 37, no. 2:
simulation of ground water flow through glacial deposits 335–349.
and crystalline bedrock in the Mirror Lake area, Grafton Winter, T.C., and D.O. Rosenberry. 1995. The interaction of
County, New Hampshire. USGS Professional Paper 1572. ground water with prairie pothole wetlands in the Cotton-
Tóth, J. 1962. A theoretical analysis of groundwater flow in small wood Lake area, east-central North Dakota, 1979–1990.
drainage basins. In Proceedings of Hydrology Symposium No. Wetlands 15, no. 3: 193–211.
3, Groundwater, 75–96. Ottawa, Canada: Queen’s Printer. Winter, T.C., and D.O. Rosenberry. 1997. Physiographic and
Winograd, I.J., and W. Thordarson. 1975. Hydrogeologic and geologic characteristics of the Shingobee River headwaters
hydrochemical framework, south-central Great Basin, area. In Hydrological and Biogeochemical Research in the
Nevada-California, with special reference to the Nevada Shingobee River Headwaters Area, North-Central Min-
Test Site. USGS Professional Paper 712-C. nesota, ed. T.C. Winter, 11–17. USGS Water-Resources
Winter, T.C. 1976. Numerical simulation analysis of the interac- Investigations Report 96–4215.
tion of lakes and ground water. USGS Professional Paper Winter, T.C., D.O. Rosenberry, D.C. Buso, and D.A. Merk.
1001. 2001. Water source to four U.S. wetlands: Implications for
Winter, T.C. 1983. The interaction of lakes with variably satu- wetland management. Wetlands 21, no. 4: 462–473.
rated porous media. Water Resources Research 19, no. 5: Wood, W.W., and W.E. Sanford. 1990. Ground-water control of
1203–1218. evaporite deposition. Economic Geology 85, 1226–1235.

1000 T.C. Winter et al. GROUND WATER 41, no. 7: 989–1000

Das könnte Ihnen auch gefallen