Integrated conduit systems discharging at springs dominate the flow

advertisement
1
2
3
4
5
6
7
8
9
10
River reversals into karst springs: a model for cave enlargement in eogenetic karst
aquifers
Jason Gulley, Jonathan B. Martin, Elizabeth J. Screaton, Paul J. Moore*
Department of Geological Sciences
P.O. Box 112120
University of Florida
Gainesville, FL 32611-2120
1
11
12
Abstract
Most conceptual models of epigenic conduit development assume that conduits
13
sourcing karst springs form as water that is undersaturated with respect to carbonate
14
minerals flows from recharge to discharge points. This process is not possible in springs
15
fed by distributed recharge that is transmitted through aquifer matrix porosity, such as
16
unconfined aquifers in eogenetic carbonate rocks. Diffusely recharged water has a long
17
residence time within the aquifer, and thus would have equilibrated with the aquifer rocks
18
prior to discharge to the conduits. The Upper Floridan aquifer (UFA) has high matrix
19
permeability (~10-13 m2) and many springs lack discrete inputs of undersaturated
20
allogenic water in their recharge areas. Consequently, another explanation for their
21
development is necessary. During flooding of the Suwannee River in north-central
22
Florida, water highly undersaturated with respect to carbonate minerals commonly
23
recharges the UFA through spring vents, and solution scallops oriented away from the
24
vents suggests most dissolution along conduit walls occurs during these flow reversals.
25
During a single flow reversal at the Peacock Spring cave system, flood water was capable
26
of dissolving up to 3.4 mm of the conduit wall rock. Dissolution occurs as flow reversals
27
follow pre-existing features that include joints and paleo-water table caves. Lack of
28
speleothems in conduits in the UFA has been used as evidence that the caves formed in
29
the phreatic zone; however, flooding would dissolve any speleothems that may have
30
formed during previous subaerial exposure. Conduit enlargement during flow reversals
31
suggests that dissolution can progress in the normal upstream directions and this process
32
may be an important driver of dissolution in any karst aquifer with outflows to surface
33
water that are subject to flooding. Flow reversals would also introduce dissolved organic
2
34
carbon and oxygen into the groundwater and provide important energy sources for cave
35
ecosystems as well as altering redox chemistry of the aquifer water.
36
Introduction
37
Conduits commonly form in diagenetically mature carbonate aquifers with low
38
matrix porosity and permeability (termed telogenetic karst by Vacher and Mylroie, 2002)
39
when undersaturated allogenic runoff flows into discrete recharge points such as
40
sinkholes or swallets. This recharge dissolves the rock along joints and bedding planes,
41
thereby expanding these preferential flow paths into conduits (Palmer, 1991). Fully
42
mature conduits thus often link recharge and discharge points in these systems.
43
In contrast, processes forming conduits remain poorly understood in aquifers with
44
high matrix porosity and permeability (termed eogenetic karst by Vacher and Mylroie,
45
2002). These aquifers typically occur in tropical marine settings and have not undergone
46
burial diagenesis that would occlude the primary depositional porosity and permeability.
47
Because the high permeability matrix allows rapid infiltration of recharge as diffuse flow
48
through the surface (e.g., Ritorto et al., 2009), point recharge (i.e. allogenic recharge) at
49
sinking streams is less common than in telogenetic karst aquifers and generally only
50
occurs where streams flow off confining layers onto the carbonate aquifer (e.g., Screaton
51
et al., 2004). Where sinking streams do exist, their potential for focused dissolution is
52
greatly diminished because of the large volumes of water stored in the matrix porosity,
53
which is commonly equilibrated with carbonate minerals of the aquifer (Moore et al.,
54
2009).
55
Little allogenic recharge occurs in the eogenetic karst of the Upper Floridan
56
aquifer (UFA), except for where streams flow off the edge of the confining layer and into
3
57
the unconfined aquifer. Nonetheless, many of Florida’s springs discharge from laterally
58
extensive phreatic conduit systems (Florea and Vacher, 2007; Martin and Gordon, 2000).
59
Because of their distance from the coast, these conduits could not have formed from
60
mixing of fresh and saline water as has been proposed for caves in the eogenetic
61
limestone of the Yucatan (Smart et al., 2006) and the Bahamas (Mylroie and Carew,
62
1990). The general lack of allogenic recharge limits input of water undersaturated with
63
respect to carbonate minerals into pre-existing high-permeability zones at the upstream
64
end of the conduits. These conduits are assumed to have formed in the phreatic zone and
65
not been subject to past subaerial exposure because most lack speleothems, unlike
66
conduits in the Yucatan and the Bahamas. Conduit dissolution has been proposed to
67
occur in phreatic zone from “headward sapping” in which high permeability zones act as
68
low resistance drains and cause flow paths to converge and concentrate dissolution,
69
further focusing flow and dissolution (c.f., Rhoades and Sinacori, 1941; White, 2001).
70
However, water from the UFA is generally saturated with respect to calcite (Martin and
71
Gordon, 2000; Moore , 2009), and thus headward sapping is unlikely to form the conduits
72
found there. Consequently, the origin of submerged eogenetic karst conduits remains
73
unresolved.
74
In this paper, we use legacy data and new observations from the Suwannee River
75
watershed in north-central Florida to suggest that reversals of springs during flood events
76
provide a mechanism to form or enlarge conduits. Spring flow reverses when river stage
77
increases faster than the hydraulic heads in the aquifer. Although backflooding of air-
78
filled caves has been observed in telogenetic karst regions (White and White, 1989) and
79
some Florida springs have been reported to reverse (Opsahl et al., 2007), the importance
4
80
of chemical processes such as dissolution during spring reversals has not been evaluated.
81
We establish that surface water is undersaturated with respect to carbonate minerals
82
during high discharge events and use observations of solution scallop direction in two
83
water-filled conduit systems to support the concept that dissolution occurs during spring
84
reversals. We collected high-resolution specific conductivity records within two conduit
85
systems and geochemical data during reversal of one spring to document the influx of
86
highly undersaturated water during flooding. These data allow an assessment of the
87
magnitude of dissolution by flood waters.
88
89
Study Locations
90
The Suwannee River watershed in north-central Florida is entirely underlain by
91
the Floridan Aquifer System (FAS), a thick sequence of limestone and dolomite that is
92
subdivided into the UFA, a middle confining unit (where it exists), and the Lower
93
Floridan aquifer (Miller, 1986). The Cody Scarp generally marks the boundary between
94
the confined and unconfined regions of the UFA and separates the Northern Highlands
95
and Gulf Coastal Lowlands physiographic areas (Fig. 1). In the Northern Highlands, the
96
UFA is overlain by the confining siliciclastic Hawthorn Group and the Surficial Aquifer
97
System. Water sources to the Suwannee River include the Surficial Aquifer System and
98
runoff, which provide tannic-rich water due to organic matter contributions from
99
wetlands. Downstream of the Cody Scarp, the Suwannee River watershed transitions to
100
being sourced by the UFA, including discharge from more than 100 springs (Rosenau et
101
al., 1977; Scott et al., 2004). These springs include 9 of Florida’s 27 first magnitude
102
springs, which are defined as having a discharge of > 2.8 m3/sec, (i.e., > 100 cfs:
5
103
Meinzer, 1927). At baseflow, these springs discharge water that is saturated with respect
104
to carbonate minerals, reflecting equilibration with the aquifer rocks.
105
Groundwater of the UFA has higher specific conductivity than surface water
106
because of the high dissolved load of carbonate minerals but it has lower dissolved
107
organic carbon concentrations, and thus does not have the characteristic tannic stain of
108
water from the surficial aquifer system or surface water draining the Northern Highlands.
109
Differences in specific conductivity and staining between groundwater and surface water
110
are particularly strong during floods (e.g., Moore et al., 2009), providing natural tracers
111
that allow separation of flood water flowing off of the Northern Highlands from
112
groundwater of the UFA.
113
Flooding is common in winter and spring from rainfall associated with cold fronts
114
and in late summer and fall from tropical storms (Grubbs and Crandall, 2007). These
115
floods frequently elevate river water levels that have their headwaters in the Northern
116
Highlands above the hydraulic head of the unconfined UFA (Martin and Dean, 2001;
117
Martin et al., 2006; Ritorto et al., 2009). The changing hydraulic gradients cause springs
118
to reverse, as shown by whirlpools at spring vents and changes in water levels and
119
tannins in wells up to 4.8 km from the river (Crandall et al., 1999).
120
We investigate flow and chemical composition of water at two springs (Madison
121
Blue and Peacock) that discharge within the Suwannee River watershed in north-central
122
Florida (Fig. 1). Madison Blue Spring is classified as a first magnitude spring that
123
contributes baseflow discharge of 2.0 to 3.9 m3/sec to the Withlacoochee River via a
124
short stream connected to the spring vent known as a spring run (Rosenau et al., 1977;
125
Scott et al., 2004). The Withlacoochee River is a major tributary to the Suwannee River,
6
126
and Madison Blue Spring is located about 12 km upstream of the convergence of the two
127
rivers (Fig. 1). More than 8 km of passages have been mapped in Madison Blue Spring
128
(Fig. 2A). Additional passages are known but have not yet been surveyed.
129
Peacock Spring, located about 67 km downstream of Madison Blue Spring and
130
2.3 km north of the Suwannee River (Fig. 1), lacks a conduit connection or a spring run
131
to the Suwannee River. The ‘spring’ is technically a group of water-filled sinkholes (karst
132
windows) that lead to 7.5 km of mapped conduits (Fig 2B). The Suwannee River
133
periodically floods and inundates the spring along a normally dry channel that connects
134
the river to the entrance of the conduit system. The channel contains a sill that restricts
135
direct infiltration of river water into conduits to times when river water elevation exceeds
136
~8 masl (Rick Owen, Florida Department of Environmental Protection, Personal
137
Communication).
138
We have also made cave-diving observations of conduit wall morphology at
139
Madison Blue and Peacock Springs, as well as at two other locations, Little River and
140
Cow Springs, that were not sampled for this study. Little River Spring is located about
141
18 km downstream from Peacock Spring and Cow Spring is located a few hundred
142
meters north of the Suwannee River about 3 km southeast of Peacock (Fig. 1).
143
144
145
Methods
Legacy flow data were collected by the Suwannee River Water Management
146
District (SRWMD) and the United States Geological Survey (USGS) for the
147
Withlacoochee River near Madison Blue Spring at Lee, Florida (USGS station 02319394)
148
and the Suwannee River near Peacock Spring near Luraville, Florida (USGS station
7
149
02320000). The Lee station is approximately 10 km downstream from Madison Blue
150
Spring and the Luraville station is approximately 3 km upstream from Peacock Springs
151
(Fig. 1). Discharge data from Madison Blue Spring were provided by the USGS from
152
continuous velocity measurements from a current meter (USGS station 02319302).
153
Chemical composition data was also collected by the SRWMD for water discharging
154
from Madison Blue Spring and from the rivers at the Lee and Luraville stations. Similar
155
flow and chemistry data are unavailable for Peacock Spring. We used the chemistry data
156
to calculate calcite saturation indices (SIcal) using PHREEQC with the LLNL database
157
(Parkhurst and Appelo, 1999). We define SI values here as the log of the ion activity
158
product divided by the equilibrium constant for calcite dissolution reaction.
159
Two observation periods of spring reversals are reported here: one in fall 2008
160
and the other in spring 2009. During fall 2008, Tropical Storm Fay passed through the
161
area causing a minor flood. This event was recorded by a Schlumberger Conductivity-
162
Temperature-Depth (CTD)-diver that was installed at the entrance to Peacock Spring to
163
make time-series measurements of specific conductivity (SpC) and temperature (T).
164
In spring 2009, specific conductivity and T were also monitored at 20-minute
165
intervals during and following major flooding in April 2009 with CTD-divers installed at
166
the entrances to Peacock and Madison Blue springs and at six locations within conduits
167
sourcing these springs (Fig. 2). CTD-divers were installed within the conduits at
168
distances from the main spring vent of 152 m (Martz Sink), 610 m (Courtyard) and 1097
169
m (Back Section) in the Madison Blue Spring conduits, and at distances of 214 m
170
(between Pothole and Olsen sinks), 884 m (Challenge Sink), and 1067 m (Distance
171
Tunnel) in the Peacock Spring conduits. CTD-divers have accuracies for T of + 0.1º C
8
172
and SpC of +1%. The CTD-divers record pressure to a maximum depth of 10 m of
173
water, which was exceeded during most of the flood and thus we have no data for water
174
depths. CTD-divers were also installed at the Lee and Luraville stations during spring
175
2009 but the flood covered both sensors with sediment and prevented any data collection.
176
To complement the SpC and T data, water was collected six times (16 and 24
177
April, 1, 8, and 15 May, and 14 July 2009) during the April 2009 flood and its recession
178
at the Luraville station and from two sinkholes, Challenge and Orange Grove sinks, that
179
intersect Peacock Spring conduits (Fig. 2B, Table 1). Samples could not be collected at
180
Madison Blue Spring during the April 2009 flood because roads to the spring were
181
submerged, making the spring inaccessible. Water was collected by extending a PVC
182
tube from the banks to directly above the center of the sinkholes and to about 5 m from
183
the banks of the river. The tubing was connected to a peristaltic pump, which drew water
184
into an overflow cup. The water was monitored for its SpC, T, dissolved oxygen (DO)
185
concentration, and pH using a calibrated YSI model 566 multi-parameter field meter, and
186
pumping continued until all values stabilized. Following stabilization, samples were
187
collected in PVC bottles for analyses of alkalinity and major element concentrations and
188
kept chilled until measurement. Samples for measurements of cation concentrations were
189
preserved with nitric acid. Alkalinity was titrated within one day of collecting the
190
samples using the Gran method (e.g., Drever, 1997) and the major element concentrations
191
were measured using a Dionex Model 500DX ion chromatograph (IC) in the Department
192
of Geological Science, University of Florida.
193
194
Most of the samples collected during the first three weeks of the flood had very
low solute concentrations and were at or near the detection limit of the IC. Consequently,
9
195
their charge balance errors are large, averaging around 17%. Charge balance errors
196
(CBEs) are less for samples collected during the flood recession, averaging around 4%.
197
We used PHREEQC (Parkhurst and Appelo, 1999) to calculate SIcal based on these data.
198
Charge balance was alternately forced on Ca2+ and alkalinity concentrations by
199
increasing or decreasing the concentration of Ca2+ and alkalinity within the calculations
200
until charge balance was achieved to assess the impact of CBEs on calculated calcite
201
saturation index. Forcing charge balances changes the SI less than 1 SI unit for samples
202
with high CBEs and for most samples changes the SI less than 0.l SI unit. While an error
203
margin of 1 SI is large, the sample with largest CBE was still had an upper calcite SI of
204
nearly -4 and reflects that the water is capable of dissolving considerable amounts of
205
calcite.
206
While some data was lost due to equipment being damaged by flooding or
207
logistical constraints, data that were collected clearly demonstrate large volumes of
208
undersaturated river water flow into springs during springs reversals. Missing data
209
include discharge at Madison Blue spring for the April 2009 flood, which destroyed the
210
USGS discharge gauging station at Madison Blue Spring. Peacock Spring is not gauged
211
because it lacks a spring run. We thus estimated the rate and volume of river water
212
intruding into the springs by dividing the distance between CTD divers by the time it
213
took for flood water to pass them, as estimated from changes in SpC of the water. We
214
assume river water flowed into the conduits during the time of decreasing and sustained
215
low conductivity and that springs began to discharge when the SpC rose following the
216
maximum flood elevation. To estimate the total amount of recharge during a reversal,
217
our calculations assume flow was maximum at the start of the reversal and decreased
10
218
linearly until the reversal stopped. We convert the flow rate to a volume of water based
219
on an estimated average conduit diameter of 3 m, which is consistent with our
220
observations of the water-filled conduits. There are large variations in conduit diameter
221
and cross-sectional morphology in both systems, 3 m is considered to be a rough average.
222
More accurate assessment of influx volumes would require detailed conduit cross section
223
measurements and either continuous velocity measurements or head data from the
224
conduits and surface water, which were not available.
225
Results
226
SICal response to elevated river discharge. Legacy data demonstrate that the
227
Withlacoochee and Suwannee rivers have an inverse exponential relationship between
228
discharge and SIcal (Fig. 3). Highest river discharges approach 500 m3/sec in the
229
Withlacoochee River at the Lee station and 1000 m3/sec in the Suwannee River at the
230
Luraville station. Water during high flow events can reach SIcal values of < -4 at both
231
stations across a wide range of discharges and most likely reflects the importance of
232
antecedent aquifer heads as a control on river water chemistry.
233
Discharge data from Madison Blue Spring reflect frequent reversals (Fig. 4).
234
Over the period of record, the volume of backflow is around 7% of discharge from
235
Madison Blue Spring. Most water sampled from the spring’s head pool is within about
236
0.2 SI units of saturation with respect to calcite. During the summer of 2007, Madison
237
Blue did not experience any reversals (Fig. 4) and SIcal was, with one exception, either
238
saturated or slightly supersaturated. The lowest values of SIcal are reported from sampling
239
trips near the end of a reversal and are only -0.6, much closer to saturation than the < -4
240
values found for the river during floods (Fig. 3A). Samples during these times probably
11
241
reflect a combination of river water and groundwater. Spring samples were not generally
242
collected during peak flood times because of limited access to the spring, so that
243
sampling times for the spring water shown in figure 4 do not correspond to sampling
244
times for the river shown in figure 3A.
245
Observational evidence for conduit enlargement during reversals. During cave
246
dives at Little River and Cow springs we observed well-developed scallops on conduit
247
walls where there are constrictions in the conduits (Fig. 5). Scallops were not observed
248
during cave dives at Peacock and Madison Blue Springs but no systematic search has
249
been made for them. In addition, the portions of passages we have observed in these
250
springs generally lack significant constrictions and conduits tend to be larger with slower
251
flow velocities than at Little River or Cow springs.
252
Response of Springs to Floods - SpCond and discharge. The passing of Tropical
253
Storm Fay resulted in the Suwannee River rising from ~5.4 m to ~7.7 meters above sea
254
level (masl) at the Luraville station, which is below the ~8 masl threshold required to
255
flood overland into the Peacock Spring karst windows. Nonetheless, we observed tannic
256
water in the conduit system and the CTD-diver installed in the spring vent recorded a
257
decrease of SpC of around 60 µS/cm during the flood (Fig. 6). This drop in SpC
258
occurred at the same time as an increase in temperature of around 0.3º C. A nearly
259
instantaneous drop in SpC from ~400 µS/cm to 357 µS/cm occurred at the time of the
260
maximum river stage. This minimum SpC value occurred approximately 6 days after the
261
maximum flood elevation in the river.
262
263
During the April 2009 flood, the stage of the Withlacoochee River at the Lee
station rose from about 9.3 to 19.5 masl in 13 days and the Suwannee River at the
12
264
Luraville station rose from about 5.7 to 14.2 m in 15 days, resulting in overland flow to
265
Peacock Spring (Fig. 7A). Recession from the flood peak required more than 6 weeks at
266
both gauges.
267
At the entrance to Peacock Spring, SpC initially increased from 400 to 414 S/cm
268
between 12:51 h and 15:51 h on 4 April as the flood reached the spring before dropping
269
to a minimum value of 24 S/cm at 23:51 h on 6 April at the entrance (Fig 7B). SpC
270
decreased rapidly to minimum values over a few hours at each CTD location with the
271
onset of the flood. Each minimum displays a small time lag related to increasing distance
272
from the cave entrance. Minimum values of SpC were lowest at the entrance (24 S/cm)
273
but were slightly higher with increasing distance from the entrance. SpC reached a
274
minimum of 46 µS/cm at 07:11 on 9 April between Pothole and Olsen sinks (214 m from
275
the entrance), 50 µS/cm at 16:20 on 9 April at Challenge Sink (884 m from the entrance)
276
and 52 S/cm at 18:01 h on 6 April at Distance Tunnel (1067 m from the entrance). The
277
rate of propagation of the low conductivity water from the entrance CTD-diver to the
278
CTD-diver farthest from the entrance (Distance Tunnel) indicates a flow velocity of
279
approximately 0.02 m/s.
280
SpC records during flood recession at Peacock suggest complex interactions
281
between pre and post flood waters in the conduit and matrix. There does not appear to be
282
any consistent relationship between the rate of SpC rebound and distance from the
283
entrance. SpC in the Distance Tunnel (1067 m from the entrance) rebounded most
284
quickly, reaching a maximum of 352 S/cm at 07:31 h on 3 May before beginning a
285
gradual decline. None of the other CTD-divers record similar maximum but they do show
286
an inflection as the rate of increase in SpC slows.
13
287
There is considerably more structure to the SpC records at Madison Blue Spring
288
than at Peacock Spring (Fig. 7A). The CTD-diver at the entrance to Madison Blue
289
Spring dropped from a background value of around 286 S/cm to a flood value of around
290
50 S/cm in 45 hours, between 00:03 h on 1 April 2009 and 21:18 h on 2 April 2009
291
(Fig. 7A). This drop in SpC occurred as the river stage increased from around 9.5 to 10.9
292
masl, elevations much below the flood peak at 19.5 masl. All of the other CTD-divers
293
record a similar rapid drop, although they lag the CTD-diver at the entrance.
294
Specific conductivity at Back Section (1097 m from the entrance) reached a
295
minimum of 47 µS/cm on 8 April, and values at Courtyard (610 m from the entrance)
296
fluctuated between around 45 and 60 µS/cm until 03:24 h on 13 April. Between this time
297
and 03:00 h on 15 April, SpC at both Courtyard and Back Section rapidly increased to
298
170 µS/cm over a period of approximately 48 hours. At Martz Sink (152 m from the
299
entrance), SpC slowly decreased to a minimum value of slightly less than 40 µS/cm on 6
300
April and then gradually increased to a value of about 85 µS/cm at 01:03 h on 17 April
301
and approximately six hours later at 07:18 h on 17 April at the entrance. After these sites
302
reached a value of around 85 µS/cm, the SpC increased rapidly to around 250 µS/cm at
303
15:50 on 17 April at Martz Sink and about six hours later at 21:18 on 17 April at the
304
Entrance. A second, smaller reversal began 18 April at the entrance and Martz Sink but
305
these small reversals do not occur farther back in the conduit system at the Courtyard or
306
Back Section CTD-divers.
307
Peacock Springs - SICal response during reversal. Chemical compositions
308
measured on samples collected from Peacock Spring represent the first systematic
309
sampling of a spring through a flood reversal (Fig. 7B). At the peak of the April flood on
14
310
16 April, approximately 9 days after low conductivity water entered the conduit system at
311
Peacock Spring, the SIcal at Orange Grove and Challenge sinks were found to be around -
312
5.0, slightly lower than the value of -4.5 found for the Suwannee River (Fig. 6B; Table
313
1). Seven days later on 24 April the SIcal had increased to –3.3 at Challenge Sink, but
314
remained low at –4.6 at Orange Grove, while this value had increased to -2.7 in the river
315
water. SIcal values at both locations increased to around -1 on 1 May, simultaneously
316
with the rapid increase in SpC at Orange Grove and Challenge sinks, while SIcal
317
estimated for the river water was slightly lower at -1.5. After 1 May, the SIcal values
318
slowly increased to around –0.5 for both Challenge and Orange Grove sinks on 14 July,
319
and the Suwannee river value approached equilibrium with calcite with a value of SIcal of
320
-0.2 (Table 1).
321
322
323
Discussion
The differences in chemical compositions of water sources to the Suwannee River
324
in the Northern Highlands and Gulf Coastal Lowlands affect the correlation between
325
discharge and SIcal (Fig. 3). The best correlation at our two sites between discharge and
326
SIcal (r2 = 0.86) exists for the Withlacoochee River near Madison Blue Spring (Fig 3A).
327
This site is located near the Cody Scarp (Fig. 1), and river water has had little opportunity
328
to react with carbonate minerals or mix with groundwater that has equilibrated with
329
carbonate minerals (e.g., Moore, 2009). In contrast, there is more scatter in the
330
correlation downstream at the Luraville station (r2 = 0.67) (Fig 3B). This station is
331
located nearly 70 km downstream of the Cody Scarp and thus changes in SIcal are
332
influenced by variable contributions from the UFA. The scatter of SIcal with discharge at
15
333
the Luraville station reflects the control of groundwater over the SIcal of the river water.
334
The fractional volumes of these two sources can vary depending on runoff and the
335
magnitude and direction of the head gradient between the aquifer and the river (e.g.,
336
Moore et al., 2009).
337
The low SIcal values calculated from legacy data, along with low SIcal values
338
estimated from samples collected during flood events in karst windows far from the river
339
(Fig. 7B), indicate that dissolution occurs as a result of the recharge of undersaturated
340
river water during spring reversals. Flow direction during times of dissolution can be
341
estimated based on the orientation of solution features such as scallops on conduit walls,
342
which form by dissolution in back-eddy currents in turbulent flow and are strongly
343
asymmetrical in the direction of flow (Bretz, 1942; Curl, 1974). The orientation of the
344
scallops at Little River (Fig. 5) and Cow springs indicate that flow was into the cave
345
during times of dissolution, as would be expected from the low SIcal values of river water
346
during flood times (Fig. 3). We have observed rapid flow into conduits as whirlpools
347
formed over the spring vent of Little River Spring during floods. The only known
348
allogenic recharge to Cow Spring is from backflooding during high river stages.
349
Dissolution would only occur during these backflooding events because matrix water in
350
the UFA is at equilibrium with calcite.
351
Mixing of flood water and pre-flood groundwater. Observations from the karst
352
window at Peacock Spring during flooding caused by Tropical Storm Fay in August and
353
September 2008 indicate river flooding can affect regions of the aquifer that do not have
354
direct conduit or overland flow connections to the river (Fig. 6). Although the river level
355
did not exceed the sill elevation at Peacock Spring, the decrease in conductivity at the
16
356
conduit entrance approximately 6 days after the flood peak indicates dilute flood water
357
flowed into the cave. The 6 day lag for the decrease in SpC is about 3 times longer than
358
the time for the decrease in conductivity observed for the April 2009 flood in which the
359
sill depth was exceeded (Fig. 7B). The greater lag time when the sill was not flooded
360
reflects slower influx of flood water, probably through secondary permeability features
361
such as joints, rather than rapid influx of surface water. Although we have no samples
362
from flooding during Tropical Storm Fay, SpC was measured to be 358 µS/cm. Because
363
groundwater in this part of the UFA is of the Ca-HCO3 type, SpC can be used as a
364
qualitative proxy for SIcal (Krawczyk and Ford 2006). We can therefore estimate calcite
365
saturation from a sample collected on 14 July 2009 at Challenge Sink which had an
366
identical SpC of 358 µS/cm and a SIcal of -0.58. Dissolution was likely if water during
367
Tropical Storm Fay had a similar undersaturation with respect to calcite.
368
The sharp drop in SpC clearly indicates that 2009 flood waters entered conduits
369
on 1 April at Madison Blue Spring and on 5 April at Peacock spring (Fig. 7); however,
370
the return of pre-flood water to the conduits (return of groundwater outflow) is difficult to
371
estimate. We interpret the rapid increase in SpC that occurred on 13 April at Back
372
Section and Courtyard, 17 April at Martz Sink and Entrance of Madison Blue Spring, and
373
on 27 April at Peacock (Fig. 7) to represent the time when mostly pre-flood ground water
374
returned to the conduits. Even after the pre-flood water enters the conduits, none of these
375
sites return to the background SpC values following the rapid rise, indicating the conduits
376
retain a fraction of flood water for at least six weeks after the flood peak (Fig 7). The
377
gradual increase in SpC and the increase in SIcal of the water prior to this time could
378
represent reactions of flood water with the aquifer rocks, mixing with pre-flood ground
17
379
water that had previously equilibrated with the aquifer rocks, or both. Floodwater that
380
had infiltrated the matrix porosity thus appears to discharge slowly to the river, resulting
381
in long periods of time to react with the aquifer material.
382
At Madison Blue Spring, changes in SpC following the 2009 flood event are
383
similar at Martz and the entrance, but the response at these two locations differs from the
384
responses at Back Section and Courtyard. These responses at Madison Blue Spring also
385
differ from responses at Peacock Spring, where changes in conductivity are similar at all
386
locations throughout the flood with only slight variations at Distance Tunnel during the
387
recession. The differences between Madison Blue Spring and Peacock Spring suggest
388
that the numerous entrances into the Peacock Spring conduit system (Fig. 2B) allow rapid
389
and complete mixing of the conduit and flood waters once overland flow is established to
390
the river. In contrast, river water infiltrates and discharges primarily through the central
391
spring vent at Madison Blue Spring, and thus discharge of the flood water is controlled
392
by branching and constriction in the conduit, as well as permeability variations in the
393
matrix rocks that receive flood waters from the conduits. The more gradual return to
394
background SpC values at Courtyard and Back Section than at Entrance or Martz Sink
395
suggests that at Courtyard and Back Section the flood water has mixed more extensively
396
with the pre-flood water in the matrix and has returned to the conduits more slowly than
397
flood water in the front section of the cave. Extensive mixing and slow return to pre-
398
flood conditions could reflect decreasing flow velocities during the flood with distance
399
from the entrance.
400
401
Estimated volumes of recharged flood waters. In the Madison Blue Spring
conduit system, the passage of the low conductivity water could be tracked at each CTD-
18
402
diver location. The drop in SpC associated with the influx of flood water occurred
403
progressively later at CTD-divers located farther from the entrance. The time lag of the
404
drop in SpC with distance into the conduits reflects flow into the conduits at a rate of
405
around 0.03 m/s, similar to the 0.02 m/s rate estimated for Peacock Spring. Using an
406
estimated average cross-sectional area of the conduits (roughly estimated from highly
407
variable conduit cross-sections observed during cave-diving) of 7 m2, the initial intrusion
408
of water into the conduits is estimated to be around 0.18 m3/s. We assume the minimum
409
value in SpC that occurred around April 8 indicates the final influx of flood water, and
410
thus the spring was reversed for 7.5 days. We use this length of time for reversals to
411
estimate that about 5.8 x 104 m3 of water flowed past the most distal CTD-diver.
412
Although we have no data for hydraulic head of the groundwater during the flood, the
413
river stage continued to increase through 10 April, about 9 days after the initial drop in
414
SpC and thus the hydraulic gradient between the river and groundwater may have
415
continued to be reversed during that time. Alternatively, the gradient may flatten or
416
reverse in the farthest part of the conduit around 8 April, when the SpC at the Back
417
Section CTD-diver begins to increase (Fig. 7A). We chose to use the SpC minimum on 8
418
April as representing the end of the reversal rather than the river stage maximum on 10
419
April.
420
In the Peacock Springs system we estimate river water was flowing into the
421
conduit for about 20.5 days during the 2009 flood based on the duration of low
422
conductivity water in the system and when SpC began its sharp increase on 21 April. The
423
rate of the propagation of the low conductivity water from the Entrance CTD-diver to the
424
CTD-diver farthest from the entrance (Challenge Sink) indicates an initial flow rate of
19
425
0.14 m3 s-1. With this flow rate, the total volume of water to infiltrate past the deepest
426
CTD was also around 1.2 x 105 m3. These estimates are likely to be minimum values
427
because they include only water that flowed through a single entrance to the point where
428
the CTD-diver is farthest from the entrance. The estimate thus neglects water that may
429
have flowed into other conduit branches or river water that may have intruded into the
430
conduits at other entrances. Furthermore, as shown by the response of SpC at Peacock
431
Spring following Hurricane Fay (Fig. 6), flow may also enter the conduits through matrix
432
porosity and fractures.
433
Estimates of dissolution rates. Samples collected during the April 2009 flood at
434
Peacock Spring allow order-of-magnitude estimates of the amount of dissolution that
435
might occur as a result of spring reversals. Chemical compositions are similar for
436
samples from all locations at the peak of the flood on 16 April (Table 1). We thus use the
437
SIcal value for the sample from Challenge Sink, which is central to the Peacock conduit
438
system, to calculate dissolution on the assumption it best represents water in the conduits.
439
If all of the undersaturated flood water reaches equilibrium with calcite, the water would
440
have dissolved 4.52 mmol/L of calcite (4.44 mmol/L and 7.03 mmol/L of calcite if
441
charge balance is forced on Ca2+ and alkalinity, respectively). This estimate is a
442
maximum because it is unlikely that all the water equilibrated with calcite of the aquifer,
443
but it does provide an estimate of the amount of dissolution that could occur from
444
flooding events. This estimated value of dissolved calcite was multiplied by the volume
445
of water calculated to flow into the conduit (1.2 x 105 m3) to estimate the maximum
446
volume of calcite dissolved during the flood. Assuming a density and a molar volume of
447
calcite of 2,710 kg m3 and 36.934 cm3 mol-1, respectively (Robie et al., 1984) and a
20
448
porosity of 30% for the Ocala Limestone (Budd and Vacher, 2004), the amount of
449
bedrock dissolved was about 28.6 m3 (28.1 m3 or 53.4 m3 if charge balance was forced on
450
Ca2+ and alkalinity, respectively).
451
The amount of wall retreat can be estimated from these estimated amounts of
452
dissolution. The conduit area is estimated to be 8.33 x 103 m2 using the surveyed distance
453
from Peacock Entrance to Challenge Sink of 884 m and an estimated average conduit
454
diameter of 3 m. Given these values, the volume of bedrock that could have been
455
dissolved during the reversal equates to a maximum wall retreat of 3.4 mm (or an average
456
value of 1.6 x 10-4 m per day of reversal). Because we lack sufficient constraints on the
457
residence time of river water in the conduits and matrix to use a kinetic model to refine
458
our wall retreat estimates, we adopt the term ‘meters per day of reversal’ to distinguish
459
our estimates from direct measurements of wall retreat rates (cf. Palmer, 1991). While the
460
amount of wall retreat will be overestimated using these boundary conditions, these
461
calculations show that the amount of retreat can be substantial. If charge balance is forced
462
on Ca2+, wall retreat calculations are not significantly reduced. If charge balance is
463
forced on alkalinity, wall retreat is calculated to be 5.34 mm (2.6 x 10-4 m per day of
464
reversal). This value is nearly two orders of magnitude higher than wall retreat rates that
465
were estimated from the nearby sink-rise system at Oleno State Park, Florida (Moore,
466
2009) and estimates of maximum wall retreat in conduits in telogenetic aquifers (3 x 10-6
467
to 3 x 10-7 m/d: Palmer, 1991). Our estimated rate of retreat is likely to be a maximum
468
since a fraction of the water would have reacted with the surfaces surrounding porosity in
469
the matrix rocks rather than the conduit walls and if the flood water does not equilibrate
470
with calcite. Alternatively, dissolution within the matrix is likely to weaken the matrix
21
471
material thus allowing physical erosion of the walls during rapid flow in floods to
472
increase wall retreat (Moore, 2009). Despite the degree of uncertainty associated with
473
these calculations, there is clearly the potential for significant amounts of dissolution
474
during spring reversals.
475
We suggest that spring reversals in eogenetic rock would dissolve more rock than
476
would occur during flood conditions at sink-rise systems. In sink-rise systems the amount
477
of dissolution that occurs during flood pulses is limited by relatively short conduit
478
residence times and because most water discharges from the spring before reacting to
479
equilibrium (e.g., Martin and Dean, 1999; Martin and Dean, 2001). In contrast, where
480
flood pulses cause spring reversals, water must flow to matrix porosity because there is
481
no spring outlet for the floodwaters. Once in the matrix, long residence times and large
482
surface areas of the matrix porosity would enhance the extent of chemical reactions,
483
increasing the amount of calcite that would dissolve (e.g., Moore, 2009). The growth of
484
conduits from reversals of springs should thus provide a powerful enlargement
485
mechanism, and is also probably more effective than the headwater-sapping hypothesis of
486
Rhoades and Sinacori (1941).
487
Origin of Florida Springs. Scallop direction and spring chemistry data support
488
the hypothesis that the springs in our study have been significantly enlarged by the input
489
of undersaturated water during spring reversals. During spring reversals, water likely
490
exploited previously-existing high permeability features, including vertical joints or caves
491
formed at older, lower water tables. Water-filled conduits have been mapped at distinct
492
levels across the Florida carbonate platform, and by their correspondence with marine
493
terraces, have been proposed to reflect the elevation of the water table at the time the
22
494
caves formed (Florea et al., 2007). We hypothesize these caves are formed from the
495
input of soil CO2 that diffuses through high matrix permeability and joints in the vadose
496
zones during times of low sea level when the water table was also low (cf. Moore, 2009).
497
As water tables increased due to late Pleistocene sea level rise, the paleo water-table
498
caves would have been flooded, and isolated water-table caves could have become
499
connected to surface water as rivers incised into joints. Once this connection was
500
established, the laterally extensive paleo-water table caves would have captured
501
undersaturated water during periods of flooding, with subsequent enlargement of high-
502
permeability zones into conduits feeding the springs. This model would also explain the
503
lack of speleothems in most water-filled caves in Florida, since undersaturated flood
504
water would dissolve any speleothems that may have been formed if the caves were
505
subaerially exposed during times of lower sea level.
506
The enlargement of pre-existing joints or caves that formed at paleo-water tables
507
by reversing river water allows phreatic caves to be created at the spring entrance without
508
discrete allogenic inputs in upstream regions. High surface area of the porous rock
509
making up the Floridan aquifer would allow extensive water-rock interactions as water is
510
discharged from the conduit into the matrix (Moore, 2009) resulting in “spongework”
511
cave morphology described for porous karst aquifers (Palmer, 1991). Dissolution during
512
spring reversals may also occur in telogenetic aquifers with low matrix porosity. Low
513
matrix porosity of the telogenetic aquifers would limit exchange of water between the
514
conduit and the matrix and flood waters may extend farther into the conduits along pre-
515
existing dissolutional voids.
516
Summary and Implications
23
517
Most conceptual models of epigenic conduit development assume that conduits
518
sourcing karst springs form by the flow of undersaturated water from recharge to
519
discharge points, a process which is not possible in springs supplied primarily by
520
distributed recharge from aquifer matrix porosity such as in unconfined eogenetic
521
aquifers. Diffusely recharged water has a long residence time within the aquifer, and thus
522
would have equilibrated with the aquifer rocks prior to discharge to the conduits. Where
523
springs are subject to flow reversals during river floods, undersaturated flood water can
524
dissolve conduits from the spring entrance. Dissolution during spring reversals enlarges
525
pre-existing void spaces such as joints and horizontal caves that were likely formed at
526
paleo-water tables and subsequently connected to the river along joints. Previously, lack
527
of speleothems was used as evidence that the underwater caves in Florida formed below
528
the water table. The model proposed here indicates that the laterally-extensive horizontal
529
conduits of many of the underwater caves in Florida may have initially formed at the
530
water table similar to present day water-table caves (Florea et al., 2007). These conduit
531
systems would have begun to function as springs when channel incision exposed a joint
532
that intersected the conduit or conduit roof collapse created a connection. The conduits
533
were later modified and enlarged by dissolution during spring reversals. We suggest that
534
spring reversals would also lead to dissolution of speleothems and thus obscure evidence
535
of water-table formation and past subaerial exposure. Laterally extensive, horizontal
536
galleries are common in water-filled caves and may reflect an origin as water-table caves.
537
Consequently, they could be used to reconstruct changes in past water-table positions in
538
response to glacioeustasy and climate change (e.g., Florea et al., 2007).
24
539
Spring reversals also have implications for aquifer contamination and
540
geochemistry. Contaminant flow into karst systems is generally considered as originating
541
as flow into sinkholes and swallets. Spring reversals would provide another mechanism,
542
besides diffuse or swallet recharge, for the injection of water with distinct chemical
543
compositions. Floodwater chemistry typically has organic carbon and oxygen
544
concentrations elevated over those of groundwater. These differences in compositions of
545
floodwater and groundwater should lead to shifts in redox conditions within the aquifer
546
from initially oxic to anoxic conditions as the organic carbon is remineralized following
547
spring reversals. These changes in redox conditions would influence diagenetic reactions
548
as well as calcite dissolution. For example, injection of organic carbon into the aquifer
549
may be an important energy source for ecosystems in the typically oligotrophic
550
environment of the aquifer. Reversals of springs would allow more time for microbes to
551
oxidize the organic carbon than would be typical in flow-through systems from sinks to
552
springs (e.g., Martin and Dean, 1999; Martin and Dean, 2001). Microbes use various
553
terminal electron acceptors such as oxygen, nitrate, and metal oxides in the oxidation of
554
organic carbon and thus these reactions should also influence nitrogen and metal
555
concentrations of the flood water. Understanding these processes will require detailed
556
time-series analyses of chemical composition of water as it flows into and from reversing
557
springs.
558
559
560
Acknowledgements
25
561
We acknowledge staff at Suwannee River Water Management District for
562
supplying much of the geochemical data used in this paper. We thank cave divers Kelly
563
Jessop, Rick Crawford, Agnes Milowka, James Toland, Bill Huth and Tom Hundley for
564
assistance with the installation of CTDs and Jim Wyatt, Jeff Hancock and Wayne Kinard
565
for additional support. We thank Ken Clizbe and Bonnie Stelzenmuller for the
566
photograph of Little River Spring. We thank Art Palmer and Dave Budd for their helpful
567
reviews of an earlier draft of the paper. Two Sisters BBQ in Mayo, Florida, is
568
acknowledged for providing key motivational support for fieldwork. We acknowledge
569
support from the National Science Foundation in the form of research grants EAR-
570
0510054 and EAR-0910794, and for a Graduate Research Fellowship to JG.
571
572
573
References
574
575
576
577
578
579
Bretz, J. H., 1942, Vadose and phreatic features of limestone caves: Journal of Geology,
v. 50, p. 675-811.
Budd, D.A. and Vacher, H.L., 2004, Matrix permeability of the confined Floridan
Aquifer, Florida, USA, Hydrogeology Journal, 12, 531-549
Crandall, C. A., Katz, B. G., and Hirtin, J. J., 1999, Hydrochemical evidence for mixing
580
of river water and groundwater during high-flow conditions, Lower Suwannee
581
River Basin, Florida, USA: Hydrogeol. J., v. 7, p. 454-467.
582
583
Curl, R. L., 1974, Deducing flow velocity in cave conduits from scallops: Bulletin of the
National Speleological Society, v. 36, p. 1-5.
26
584
585
586
Drever, J. I., 1997, The Geochemistry of Natural Waters: Englewood Cliffs, NJ, Prentice
Hall, 437 p.
Dreybrodt, W., Romanov, D., and Kaufmann, G., 2009, Evolution of isolated caves in
587
porous limestone by mixing of phreatic water and surface water at the water table
588
of unconfined aquifers: A model approach: Journal of Hydrology, v. 376, no. 1-2,
589
p. 200-208.
590
591
Florea, L. J., and Vacher, H. L., 2007, Eogenetic karst hydrology: Insights from the 2004
hurricanes, peninsular Florida: Ground Water, v. 45, p. 439-446.
592
Florea, L. J., Vacher, H. L., Donahue, B., and Naar, D., 2007, Quaternary cave levels in
593
peninsular Florida: Quaternary Science Reviews, v. 26, no. 9-10, p. 1344-1361.
594
Grubbs, J. W., and Crandall, C. A., 2007, Exchanges of water between the Upper floridan
595
aquifer and the Lower Suwannee and Lower Santa Fe Rivers, Florida: U.S.
596
Geological Survey Professional Paper 1656-C, p. 83.
597
Krawczyk, W. E. and Ford, D. C., 2006, Correlating specific conductivity with total
598
hardness in limestone and dolomite karst waters: Earth Surface Processes and
599
Landforms, v. 31, p. 221-234.
600
Martin, J. B., and Dean, R. W., 1999, Temperature as a natural tracer of short residence
601
times for ground water in karst aquifers, in Palmer, A. N., Palmer, M. V., and
602
Sasowsky, I. D., eds., Karst Modeling, Karst Waters Institute Special Publication
603
#5, p. 236-242.
604
605
Martin, J.B., and Dean, R.W., 2001, Exchange of water between conduits and matrix in
the Floridan Aquifer: Chemical Geology, v. 179, p. 145-165.
27
606
Martin, J. B., and Gordon, S. L., 2000, Surface and ground water mixing, flow paths, and
607
temporal variations in chemical compositions of karst springs, in Sasowsky, I. D.,
608
and Wicks, C., eds., Groundwater Flow and Contaminant Transport in Carbonate
609
Aquifers: Rotterdam, A.A. Balkema, p. 65-92.
610
Martin, J. M., Screaton, E. J., and Martin, J. B., 2006, Monitoring well responses to karst
611
conduit head fluctuations: Implications for fluid exchange and matrix
612
transmissivity in the Floridan aquifer, in Wicks, C., and Harmon, R., eds.,
613
Geological Society of America, Special Paper 404, p. 209-218.
614
Moore, P. J., 2009, Controls on the Generation of Secondary Porosity in Eogenetic Karst:
615
Examples from San Salvador Island, Bahamas and North-central Florida:
616
University of Florida, 140 p.
617
Moore, P. J., Martin, J. B., and Screaton, E. J., 2009, Geochemical and statistical
618
evidence of recharge, mixing, and controls on spring discharge in an eogenetic
619
karst aquifer: Journal of Hydrology, v. 376, no. 3-4, p. 443-455.
620
Mylroie, J. E., and Carew, J. L., 1990, The flank margin model for dissolution cave
621
development in carbonate platforms: Earth Surface Processes and Landforms, v.
622
15, p. 413-424.
623
Opsahl, S. P., Happel, J. D., and Chanton, J. P., 2007, Unusual chemistry and
624
anthropogenic contaminants in Upper Floridan Aquifer groundwater underneath
625
the Chickasawhatchee Swamp, in Proceedings of the 2007 Georgia Water
626
Resources Conference, Athens, Georgia, p. 4.
627
628
Palmer, A. N., 1991, Origin and morphology of limestone caves: Geological Society of
America Bulletin, v. 103, p. 1-21.
28
629
630
631
Parkhurst, D. L., and Appelo, C. A. J., 1999, User's guide to PHREEQC (Version 2):
USGS.
Plummer, L. N., 1975, Mixing of sea water with calcium carbonate ground water, in
632
Whitten, E. H. T. e., ed., Quantitative Studies in the Geological Sciences,
633
Geological Society of America, Memoir 142, p. 219-236.
634
635
636
Rhoades, R., and Sinacori, M. N., 1941, Pattern of groundwater flow and solution:
Journal of Geology, v. 49, p. 785-794.
Ritorto, M., Screaton, E. J., Martin, J. B., and Moore, P. J., 2009, Relative importance
637
and chemical effects of diffuse and focused recharge in an eogenetic karst aquifer:
638
an example from the unconfined upper Floridan aquifer, USA: Hydrogeol. J.
639
Robie, R. A., Hemingway, B. S., and Fisher, J. R., 1984, Thermodynamic Properties of
640
minerals and Related Substances at 298.15 K and 1 Bar (105 Pascals) Pressure
641
and at Higher Temperatures: Washington, United States Government Printing
642
Office, 456 p.
643
Rosenau, J. C., Faulkner, G. L., Charles W. Hendry, J., and Hull, R. W., 1977, The
644
Springs of Florida: Tallahassee, Florida, Department of Natural Resources,
645
Bulletin #31, 461 p.
646
Scott, T. M., Means, G. H., Meegan, R. P., Means, R. C., Upchurch, S. B., Copeland, R.
647
E., Jones, J., Roberts, T., and Willet, A., 2004, Springs of Florida: Tallahassee,
648
Florida, Florida Geological Survey, p. 377.
649
Screaton, E., Martin, J. B., Ginn, B., and Smith, L., 2004, Conduit properties and
650
karstification in the Santa Fe River Sink-Rise System of the Floridan Aquifer:
651
Ground Water, v. 42, p. 338-346.
29
652
653
654
Vacher, H. L., and Mylroie, J. E., 2002, Eogenetic karst from the perspective of an
equivalent porous medium: Carbonates and Evaporites, v. 17, p. 182-196.
Whitaker, F. F., and Smart, P. L., 2007, Geochemistry of meteoric diagenesis in
655
carbonate islands of the northern Bahamas: 2. Geochemical modelling and
656
budgeting of diagenesis: Hydrological Processes, v. 21, p. 967-982.
657
White, W. B., 2001, Ground water flow in karst: Matrix flow and conduit flow with
658
implications for the Floridan aquifer, in Martin, J. B., Wicks, C. M., and
659
Sasowsky, I. D., eds., Hydrogeology and Biology of Post-Paleozoic Carbonate
660
Aquifers: Charles Town, WV, Karst Waters Institute, p. 9-13.
661
662
White, W. B., and White, E. L., 1989, Karst Hydrology: Concepts from the Mammoth
Cave Area: New York, Van Norstrand Reinhold, 346 p.
663
664
665
666
667
668
669
670
671
672
673
674
30
675
676
31
Download