+ All Categories
Home > Documents > To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs...

To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs...

Date post: 05-Jul-2020
Category:
Upload: others
View: 0 times
Download: 0 times
Share this document with a friend
117
ENVIRONMENTAL CONTROLS ON CARBONATE MINERAL DISSOLUTION: RATES AND MAGNITUDES By JOHN E. EZELL A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY UNIVERSITY OF FLORIDA 2016
Transcript
Page 1: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

ENVIRONMENTAL CONTROLS ON CARBONATE MINERAL DISSOLUTION: RATES AND MAGNITUDES

By

JOHN E. EZELL

A DISSERTATION PRESENTED TO THE GRADUATE SCHOOL OF THE UNIVERSITY OF FLORIDA IN PARTIAL FULFILLMENT

OF THE REQUIREMENTS FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

UNIVERSITY OF FLORIDA

2016

Page 2: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

© 2016 John E. Ezell

Page 3: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

To My Family

Page 4: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

4

ACKNOWLEDGMENTS

I would like to thank my advisor, committee, family, friends, and all those who

remind us that when you’re all alone and there’s nobody home it’s nice to be able to

count on a friend.

Page 5: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

5

TABLE OF CONTENTS page

ACKNOWLEDGMENTS .................................................................................................. 4

LIST OF TABLES ............................................................................................................ 7

LIST OF FIGURES .......................................................................................................... 8

ABSTRACT ................................................................................................................... 10

CHAPTER

1 INTRODUCTORY REMARKS ................................................................................ 12

2 TEMPORAL VARIATIONS IN CALCITE DISSOLTUION RATES DURING FLOODING OF KARST SPRINGS ......................................................................... 20

2.1 Introduction ....................................................................................................... 20 2.2 Study Sites and Methods .................................................................................. 23

2.2.1 Study Sites .............................................................................................. 23 2.2.2 Methods ................................................................................................... 25

2.2.2.1 Field methods and sample analyses .............................................. 25 2.2.2.2 Mixing estimates of river water and groundwater ........................... 27 2.2.2.3 Dissolution rates estimates ............................................................ 28 2.2.2.4 Impacts of mixing/dissolution Ca2+ additions and DOC

remineralization on carbonate dissolution ............................................... 30 2.3 Results ............................................................................................................. 31

2.3.1 Peacock Springs ..................................................................................... 31 2.3.2 Madison Blue Spring .............................................................................. 32

2.4 Discussion ....................................................................................................... 34 2.4.1 Temporal Variations in Dissolution Rates ................................................ 34 2.4.2 Comparisons of Rate Models .................................................................. 35 2.4.3 Saturation States and Dissolution Magnitude Through the Recession .... 36

2.5 Conclusions ...................................................................................................... 38

3 RELATIVE AMOUNTS OF DISSOLUTION IN CARBONATE TERRAINS FROM LOSING RIVERS AND DIRECT PRECIPITATION ................................................. 49

3.1 Introduction ....................................................................................................... 49 3.2 Study Sites and Methods .................................................................................. 51

3.2.1 Study Sites .............................................................................................. 51 3.2.2 Methods ................................................................................................... 53

3.2.2.1 Monitoring data and historical reversal estimates .......................... 53 3.2.2.2 Field methods and sample analyses .............................................. 54 3.2.2.3 Recharge and dissolution modeling ............................................... 56

3.3 Results .............................................................................................................. 60

Page 6: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

6

3.3.1 Rainfall and Recharge ............................................................................. 60 3.3.2 Hydrologic Responses ............................................................................. 60 3.3.3 Historic Record of High Water Events ..................................................... 60 3.3.4 Dissolution Estimates .............................................................................. 62

3.4 Discussion ........................................................................................................ 63 3.4.1 Dissolution Estimates .............................................................................. 64 3.4.2 Implications for Regional Geomorphology and Hydrology ....................... 67

3.5 Conclusions ...................................................................................................... 68

4 THE IMPACT OF BIOGEOCHEMICAL DRIVEN CARBONATE DISSOLUTION WITH POTENTIAL CLIMATIC IMPLICATIONS ...................................................... 76

4.1 Introduction ....................................................................................................... 76 4.2 Study Site and Methods .................................................................................... 80

4.2.1 Study Site ................................................................................................ 80 4.2.2 Field Methods and Sample Analyses ...................................................... 80 4.2.3 Estimates of Potential Dissolution and Chemical Species ....................... 83

4.3 Results .............................................................................................................. 84 4.4 Discussion ........................................................................................................ 86

4.4.1 Relative Impact of Tidal and Solar Radiation Cycles on Dissolution Potential ........................................................................................................ 86

4.4.2 Implications for Platform Dissolution........................................................ 89 4.4.3 Quantifying Dissolution Driven by Carbonic vrs. Sulfuric Acid and C

Flux to the Atmosphere ................................................................................. 90 4.5 Conclusions ...................................................................................................... 94

5 SUMMARY ........................................................................................................... 105

LIST OF REFERENCES ............................................................................................. 108

BIOGRAPHICAL SKETCH .......................................................................................... 117

Page 7: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

7

LIST OF TABLES

Table page 2-1 Various dissolution rates. ................................................................................... 40

2-2 DOC concentrations, SI values, and dissolution rates for the four sampled floods. ................................................................................................................. 41

4-1 Water chemistry and parameters for grab samples ............................................ 96

Page 8: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

8

LIST OF FIGURES

Figure page 2-1 Maps of river and cave systems. ........................................................................ 42

2-2 Conductivity and temperature through time at spring vents ................................ 43

2-3 Concentrations of calcium and calcite saturation indices through time. .............. 44

2-4 Time variations in the river stage, dissolution rates for reversals, and cumulative dissolution ........................................................................................ 45

2-5 Total calcite dissolution as measured by Ca2+ for each event by the SD equation (black) and the PWP equation (Gray). ................................................. 46

2-6 Predicted calcite dissolution values and SI range durations. .............................. 47

2-7 Measured and mixed Ca2+ concentrations and measured SI 0 and mixed SI 0 Ca2+ concentrations through time.. ..................................................................... 48

3-1 Karst terrain under normal and storm conditions.. .............................................. 70

3-2 Maps of the drainage basin and cave. ................................................................ 71

3-3 Stage data for Pinetta (dashed red line) and Madison Blue (black solid line) ..... 72

3-4 River and spring responses to rainfall.. .............................................................. 73

3-5 Limestone river bank along the Suwannee River in north Florida. ..................... 74

3-6 Long term Pinetta Gauge Station stage data from 1932 to 2013 ........................ 75

4-1 Map of The Bahamas and San Salvador.. .......................................................... 97

4-2 Conceptual model of a cross section of Inkwell Bluehole during high and low tide. . .................................................................................................................. 98

4-3 Depth profiles collected on May 6, 2012 showing............................................... 99

4-4 Dissolved oxygen, water depth, and pH plotted through time at the nearshore site in Inkwell. . ................................................................................................ 100

4-5 pH values and variation from mean water level ................................................ 101

4-6 Sulfide concentrations and variation from mean water level ............................. 102

4-7 d 13C values and variation from mean water level. ............................................ 103

Page 9: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

9

4-8 Dissolved inorganic carbon concentrations and variation from mean water level .................................................................................................................. 104

Page 10: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

10

Abstract of Dissertation Presented to the Graduate School of the University of Florida in Partial Fulfillment of the Requirements for the Degree of Doctor of Philosophy

ENVIRONMENTAL CONTROLS ON CARBONATE MINERAL DISSOLUTION: RATES

AND MAGNITUDES

By

John E. Ezell

May 2016

Chair: Jonathan Martin Major: Geology

Carbonate minerals, which are highly soluble and the largest C reservoir on

Earth, compose around 20% of Earth’s land surface. Their dissolution affects

geomorphology, hydrology, water chemical composition, and C cycling. Controls on

and rates of dissolution are therefore critical processes in carbonate terrains, which are

investigated here in northern Florida springsheds and a bluehole on San Salvador

Island, Bahamas. Carbonate dissolution results from rainfall reacting with rocks at the

land surface and forcing river water into river banks and spring systems. During one

storm in north Florida, calcite dissolution from rainfall (~6 x 109mmol) was greater than

dissolution from river water penetrating river banks (~2.4 x 109mmol) and a spring (~3.8

x 108mmol). Dissolution from rainfall occurs more frequently than dissolution from river

water; however, river intrusion focuses dissolution in small areas. Dissolution rates

increase up to 17 orders of magnitude during spring reversals resulting in episodic

dissolution in spring systems. During spring reversals, undersaturated river water

dissolves calcite and undersaturation is enhanced by organic carbon remineralization.

Dissolution rates during a spring reversal are slowed more by mixing with groundwater

near equilibrium with calcite than by addition of calcium from dissolution. Both rainfall

Page 11: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

11

and river water dissolution formed the hydrology and geomorphology of north Florida,

which includes extensive cave systems and surface rivers. In blueholes, tidal and diel

cycles control water-column chemistry through photosynthesis/respiration and water

column variations. In Inkwell bluehole, up to 0.8mmol/L of sulfuric acid is formed during

the day, decreasing water column pH by up to 0.31 pH units, and up to 0.4mmol/L of

carbonic acid is formed at night, reducing water column pH by as much as 0.18 units.

Calcite dissolution potential increases by 0.62mmol/L from sulfuric acid but only

0.25mmol/L from carbonic acid. Since sulfuric acid dissolution provides CO2 while

carbonic acid dissolution removes CO2 from the atmosphere, the greater dissolution

from sulfuric over carbonic acid produces a net source of carbon to the atmosphere.

These results show that carbonate dissolution’s impact on landscapes, regional

hydrology, and global carbon cycle depends on causes, rates, and distribution of

dissolution, which are variable across carbonate landscapes.

Page 12: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

12

CHAPTER 1 INTRODUCTORY REMARKS

Carbon is an integral building block of life for all flora and fauna and is stored in

in three distinct natural reserves: the atmosphere, the lithosphere, and the hydrosphere.

Shifts from one reserve to another can impact water resources, geomorphology, and

Earth’s climate. Transitions between the reserves are constantly occurring and have the

potential to impact life on Earth, but these transitions remain poorly understood in many

settings.

There is ~7 x 107 Pg C stored in carbonate rocks around the world (Martin et al.,

2013), which represents the largest reserve on Earth. This reserve fluctuates over time,

gaining during carbonate precipitation and losing during carbonate dissolution.

Carbonate precipitation primarily occurs in the ocean, but dissolution can occur

anywhere there is carbonate and fluid undersaturated with respect to carbonate

minerals. We therefore focus on carbonate dissolution in this work.

The total volume of carbonate that will dissolve is dependent on the duration of

carbonate dissolution and the rate at which the carbonate is dissolving, since natural,

flowing fresh waters frequently do not reach complete equilibrium with respect to

carbonate minerals. Carbonate dissolution rates are controlled by the undersaturation of

the water and the ability of water to transport reaction byproducts away from the

reaction site. Dissolution rates are not linear across a range of undersaturations

(Dreybrodt, 1990; Dreybrodt, W., 1998) Rates slow rapidly when ~30% saturation is

reached (Kaufmann and Dreybrodt, 2007) and again somewhere between 60% and

85% saturation depending on experiment construction (Svensson and Dreybrodt, 1992;

White, 2002). Another factor affecting the saturation of water with respect to a

Page 13: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

13

carbonate mineral is fluid flow in the system. If flow is sufficient to remove dissolved

ions from the reaction site, dissolution rates will be determined by the rate of CO2

conversion to carbonic acid, in most systems. If flow does not remove ions, waters

immediately next to the reaction site will become more saturated and dissolution will

slow once ion removal is dependent on diffusion rather than advection (Liu and

Dreybrodt, 1997).

The total volume of a carbonate mineral (e.g. calcite) a given amount of water

can dissolve is also dependent on the undersaturation of the water. Waters at complete

equilibrium with respect to calcite have a saturation index (SI) equal to zero and cannot

dissolve any more material. This saturation index can be lowered if the water reacts to

form acid or if the water mixes with a second water with different chemistry. Addition of

new ions to the solution can allow further dissolution as a consequence of the common

ion effect. In natural systems, fresh waters seldom reach complete equilibrium with

respect to calcite because dissolution rates slow as the water approaches equilibrium

with respect to calcite.

Natural controls on the undersaturation of water with respect to a given

carbonate mineral usually include the quantity and type of acids produced and how

much of a mineral the water has already dissolved. Carbonate dissolution is frequently

driven by carbonic acid, which is formed when rainfall reacts with CO2 produced by the

remineralization of organic carbon, atmospheric CO2, and/or soil CO2 (Eq. 1-1).

Carbonic acid reacts with the basic carbonate minerals, in this case limestone, and a

Ca2+ ion and two bicarbonates are released (Eq. 1-2). Dissolution can also be driven by

sulfuric acid, which is formed when H2S is oxidized (Eq. 1-3). This process frequently

Page 14: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

14

occurs near the fresh water-salt water interface, where O2 and S2- are both present.

During sulfuric acid dissolution, Ca2+ and CO2 are produced along with water and

sulfate (Eq. 1-4).

𝐶𝑂2 + 𝐻2𝑂 ↔ 𝐻2𝐶𝑂3 (1-1)

𝐻2𝐶𝑂3 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 2𝐻𝐶𝑂3− (1-2)

𝐻2𝑆 + 2𝑂2 ↔ 𝐻2𝑆𝑂4 (1-3)

𝐻2𝑆𝑂4 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 𝐶𝑂2 + 𝐻2𝑂 + 𝑆𝑂42− (1-4)

Each acid increases the amount of carbonate mineral (e.g. calcite) that water can

dissolve, but when the water dissolves calcite, the water comes closer to saturation with

respect to calcite. The total volume of carbonate a water can dissolve (SI = 0) is easily

determined by basic geochemical modeling of the ions and properties (e.g. temperature,

pH, etc.) of the water, but when ion concentrations and water properties change in

natural settings because of dissolution, organic matter remineralization, or mixing with a

second water source, modeling the dissolution capability of a given water becomes

much more difficult. In some surface water-groundwater interactions, carbonate

dissolution volumes and rates have yet to be estimated accurately.

Carbonate dissolution rates and volumes are important to understand and the

removal of these carbonate minerals also has implications. Dissolution begins when

carbonate platforms are exposed by uplift or falling sea levels. Carbonic acid is formed

as rainfall equilibrates with CO2 in the atmosphere and gains further CO2 from soil

microbial respiration. This acidic water dissolves soluble rock at the land surface and

may be channeled along geologic planes of weakness (joints or fractures) or flow paths

that become preferred by capturing increasing volumes of rainfall (Hanna and Rajaram,

Page 15: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

15

1998). The rainfall also picks up organic carbon from the soil. When rainwater

penetrates the platform down to the water table, the organic carbon is remineralized and

more CO2 is produced. Over time, preferential flow paths widen and deliver more

organic-carbon-rich water to the water table, where enlarged voids are created (Florea

et al. 2007, Gulley et al. 2013). These flow paths continue to transport water more

rapidly than the surrounding matrix and are enlarged by the higher flow rate. Eventually,

these voids become major conduits, which make up the traversable cave and spring

systems seen today. These conduits transport most of the subsurface flow in the Upper

Floridan aquifer hydrologic system.

Some conduits in north Florida have surface expressions as springs, which

discharge to nearby rivers. When rivers flood, river water can be forced into the aquifer

through river banks, and the river water can also reverse the flow direction of springs,

forcing river water into the spring that normally discharges groundwater to the river.

The river water is undersaturated with respect to calcite and dissolves the matrix rock,

but the extent of dissolution is not fully understood because of variations in natural

controls on dissolution that govern dissolution rates and volumes.

After terrestrial carbonate is dissolved, the dissolution byproducts enter either the

hydrosphere or atmosphere. To understand the impact of this transition from one

sphere to another, we consider dissolution/weathering of carbonate minerals on a global

scale. The two major types of weathering that occur on the planet are silicate and

carbonate weathering. Silicates are more plentiful than carbonates and weathering

silicates acts as a sink for atmospheric C (Brady, 1991). Though silicate weathering

Page 16: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

16

takes longer than carbonate weathering, silicate weathering is frequently considered the

only weathering to impact the global C cycle:

𝐶𝑂2 + 𝐶𝑎𝑆𝑖𝑂3 ↔ 𝐶𝑎𝐶𝑂3 + 𝑆𝑖𝑂2 (1-5)

Recent research showed that that carbonate weathering can impact the global C

cycle (Liu et al. 2011; Torres et al. 2014), but even as a subset of carbonate weathering,

some believe that there is no net flux of C into or out of the atmosphere when

dissolution is driven by carbonic acid (Berner et al. 1983). This belief stems from

carbonic acid dissolution reactions (Eq. 1-1 and 1-2), which show that for every C that is

sequestered from atmospheric CO2, another C is released as bicarbonate. However,

studies have shown carbonate dissolution to potentially affect the global carbon cycle

over a few hundred years (Liu et al. 2011). Liu et al. (2011) proposed that dissolved

bicarbonate was taken up from streams by plants, incorporated by the plants, washed

out to sea, and buried, thereby sequestering the C and leading to a net carbon sink in

the global C cycle.

The global C cycle can also be impacted by carbonate dissolution driven by

sulfuric acid, which results in the release of CO2 to the atmosphere, making it a net

source of C to the atmosphere (Torres et al. 2014). Therefore, carbonate dissolution

can act as a sink or source of C to the atmosphere. These sinks and sources are

important because CO2 is considered a greenhouse gas that is linked to warming and

cooling phases of Earth’s history.

In addition to comprising the largest reserve of C on Earth, carbonate rocks also

represent ~20% of the Earth’s land surface (Ford & Williams 2007). Measures of

carbonate dissolution and associated C fluctuations in the various reserves are critical

Page 17: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

17

to understanding the global C cycle budget. This research addresses carbonate

dissolution rates and quantities, as well as the resultant fluxes of C from this dissolution.

This research was conducted in north Florida and in San Salvador, The Bahamas,

enabling study of carbonate dissolution in both sub-tropical and tropical climates and in

fresh-water and brackish-water systems. We aim to better understand natural controls

on undersaturation, changes in aqueous chemistry through time, and the resulting

impacts of carbonate dissolution on the different carbon reserves.

Chapter 2 addresses carbonate dissolution rates in Peacock and Madison Blue

Spring, Florida, during four spring reversals that occurred from 2009 to 2012.

Dissolution rates declined following the start of the reversal, despite continued

production of CO2 from organic carbon remineralization. Dissolution rates of intruding

river water were slowed primarily by mixing with matrix water that was nearly saturated

with respect to calcite and secondarily by Ca2+ concentrations generated by river water

dissolution of calcite. Conduit expansion is primarily driven by dissolution that occurs

during spring reversals.

Chapter 3 re-examines one of the spring reversals studied in Chapter 2, which

occurred at Madison Blue Spring in 2012. This work estimated total dissolution that

occurred during the spring reversal, along the river, and across the drainage basin.

River water drives dissolution by triggering a spring reversal and penetrating the river

banks to reach and react with the rock matrix, but rainfall drives dissolution across the

drainage basin by filtering through the soil and reacting with the rock matrix. These

dissolution estimates were compared to determine which process was most responsible

for shaping the landscape seen today. This work used time series geochemical data

Page 18: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

18

and hydrologic modeling to generate the first quantitative estimates of dissolution that

occurred during a spring reversal. Overall, rainfall dissolved more calcite than the spring

reversal or river bank penetration, but rainfall dissolution was spread over the drainage

basin whereas the river-water dissolution was spatially concentrated. These dissolution

patterns likely result in the current geomorphology and hydrology seen today in north

Florida. If denudation rates were faster, cave systems would be exposed before they

developed kilometers of subterranean passages, and if river-water dissolution were

greater, surface rivers would cease to exist.

Chapter 4 studies biogeochemical and tidal processes that affect carbonate

dissolution in a bluehole on San Salvador, The Bahamas. This work tracked aqueous

geochemistry changes though time and tidal cycles to determine which factors control

calcite saturation in the water column. We also investigated the respective roles of

carbonic and sulfuric acid in carbonate dissolution. Carbonic acid was found to primarily

form during the night when plants respired CO2, and sulfuric acid formed primarily

during the day when photosynthesis produced oxygen in the water column, which

reacted with H2S. The decreases in pH associated with each acid and time of day were

modeled to estimate potential carbonate dissolution. Sulfuric acid was found to dissolve

more carbonate than carbonic acid, and this bluehole study could potentially serve as

an example of processes that occur in the matrix across the carbonate platform.

These three studies all focused on carbonate dissolution, which has an important

role in regional hydrology, local geomorphology, and global carbon cycling. A better

understanding of the natural controls on dissolution rates and volumes through time is

Page 19: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

19

required to predict future impacts of carbonate dissolution on water resources,

landscape evolution, and atmospheric chemistry.

Page 20: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

20

CHAPTER 2 TEMPORAL VARIATIONS IN CALCITE DISSOLTUION RATES DURING FLOODING

OF KARST SPRINGS

2.1 Introduction

Dissolution is the defining feature of carbonate bedrock terrains, which dissolve

when they come into contact with water that is undersaturated with respect to calcite.

Dissolution increases permeability, thereby forming productive aquifers that provide

potable water to approximately 20% of the world’s population (Ford and Williams, 2007).

Dissolution commonly results from the presence of carbonic acid, formed by the

hydration of dissolved CO2:

𝐶𝑂2 + 𝐻2𝑂 ↔ 𝐻2𝐶𝑂3 (2-1)

𝐻2𝐶𝑂3 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 2𝐻𝐶𝑂3−

(2-2)

Although other acids can dissolve carbonate minerals, for example sulfuric acid

during the oxidation of dissolved sulfide (Botrell et al., 1991; Hercod et al., 1998;

Johnson and Hallberg, 2005; Spence and Telmer, 2005), reaction 2-2 represents the

most common dissolution mechanism for fresh water in the near surface (Palmer, 1991;

White, 1988; Dreybrodt and Gabrovsek, 2002). Dissolution can be a time-varying

process that depends on environmental conditions that control the composition of water,

the saturation state with respect to bedrock minerals, and rate of dissolution. Although

dissolution rates have been measured in laboratory experiments (Plummer et al., 1978;

Buhmann and Dreybrodt, 1985a, b) and time variations in the saturation states of water

in karst aquifers have been measured in natural systems (Shuster and White, 1971;

Jacobson and Langmuir, 1974), little is known about temporal variations in dissolution

rates of natural karst aquifers.

Page 21: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

21

Several experimental approaches have been used to quantify calcite dissolution

rates (Table 2-1). One early approach was based on the summation of reaction rates for

each step of the calcite dissolution reaction, including CO2 dissolution and hydration of

CO2 to carbonic acid (eq. 2-1), and subsequent dissolution of carbonate minerals

(reaction 2-2; Plummer et al., 1978). This approach uses temperature-dependent

laboratory-derived reaction time constants and activities of H+, H2CO3, H2O, Ca2+, and

HCO3- to estimate dissolution rates across a range of pH and pCO2 values. Additional

experiments have been conducted in open conditions at constant pCO2 values similar to

Earth’s atmosphere (~0.005 atm) and under closed conditions, with an initial pCO2

range of 0.01-0.1 atm, similar to what might be found in soils, to assess the role of pCO2

in calcite dissolution (Buhmann and Dreybrodt, 1985a; b). In both open and closed

systems, dissolution rates were found to depend on Ca2+ and CO2 concentrations,

temperature, and hydrodynamics of the system (Buhmann and Dreybrodt, 1985a;b)

(Table 2-1). In more recent studies, atomic force microscopy has been used to track

dissolution on a microscopic scale at atmospheric and lower CO2 concentrations (Dove

and Platt, 1996; Shiraki et al., 2000) (Table 2-1).

Most studies of dissolution rates in natural systems are commonly reported as

long-term averages of denudation rates (Table 2-1), which represent the average rate of

land surface lowering (Opdyke et al., 1984; Jennings, 1985; Adams et al., 2010). Rates

of 0.015 to 0.040 mm/yr were found by measuring the height of pedestals formed by

dissolution surrounding glacial erratics that armor soluble rocks (Jennings, 1985).

Denudation rates of up to 0.03 mm/yr (Opdyke et al., 1984) and 0.09 mm/yr (Pitty,

1968) were estimated from concentrations of dissolved ions in rivers and springs in

Page 22: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

22

Florida and the United Kingdom, respectively, assuming dissolution is the sole control

on their concentrations (Table 2-1). Adams et al. (2010) combined modeling of the

isostatic rebound of northern Florida, variations in sea level, and a relationship between

rainfall and dissolution to estimate denudation rates in Florida of around 0.09 mm/yr

over the past 2 million years, focusing on the same region studied by Opdyke et al.

(1984) (Table 2-1). Adams et al. (2010) estimated rates that were about 3.5 times

faster than those found by Opdyke et al. (1984), assuming that uplift was caused solely

by isostatic rebound.

Dissolution rates that occur within active hydrological systems, including rivers

and caves, would be expected to be much faster than whole-landscape denudation

rates because CO2 fluxes are concentrated in these zones (Covington et al., 2015).

Maintaining undersaturation requires water flow, which delivers undersaturated water to

reaction faces and removes reaction products. As undersaturation changes through

time, dissolution rates should also change. Time series measurements of aqueous

geochemistry in these zones, as reported here, could thus provide information on time

variation of dissolution rates, information that is not available from landscape

denudation studies. Because flow rates increase and saturation states decrease most

during floods, these events should dominate dissolution in karst aquifers. Floods may

be particularly important, as they introduce river water into aquifers when river

elevations rise above the hydraulic head of springs, causing flow to reverse, thereby

flooding the aquifers (Gèze, 1987; Albéric, 1998; Gulley et al., 2011),

Changes in flow direction at springs and within karst conduits dictate that

estimates of dissolution magnitude must consider time-varying dissolution rates. Little is

Page 23: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

23

known of the magnitude of changes in dissolution rates between flood and baseflow.

Even less is known about the specific processes that cause dissolution rates to change,

which include calcite dissolution, mixing between intruded water and groundwater

during recession, and reactions such as DOC remineralization. Consequently, in this

paper we: (1) calculate rates and magnitudes of calcite dissolution as river water

intrudes into the upper Floridan aquifer using its chemical composition and two

laboratory-based models with differing structures and assumptions, and (2) evaluate

potential controls on changing dissolution rates. Our calculations of the magnitude of

dissolution refine previous estimates made by Gulley et al. (2011) for one intrusion

event, and include estimates of dissolution rates and magnitudes of three additional

intrusion events characterized by a range of magnitudes.

2.2 Study Sites and Methods

2.2.1 Study Sites

The two spring systems in this study, Madison Blue and Peacock Springs, are

located in north-central Florida (Fig. 2-1). Madison Blue Spring discharges to the

Withlacoochee River ~20 km upstream of its confluence with the Suwannee River.

Madison Blue Spring is located near the Cody Scarp, which is a regional geomorphic

feature that marks the erosional edge of the Miocene Hawthorn Group (Fig. 2-1A). The

Hawthorn Group rocks are predominately low-permeability siliciclastic sands and

claystones that confine the Upper Floridan aquifer to the north and east of both spring

systems. Peacock Springs is located on the Suwannee River ~55 km downstream of

the confluence with the Withlacoochee River. The Withlacoochee and Suwannee

Rivers are both tannic and primarily receive runoff from wetlands, agricultural fields, and

forested land. The low permeability of the Hawthorn Group can produce large volumes

Page 24: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

24

of surface runoff that cause common floods on the two rivers and intrusion of river water

into the springs (Gully et al., 2011; Brown et al., 2014).

Both Madison Blue and Peacock springs have extensive mapped conduit

systems and exploration is on-going. Conduits for both spring systems occur in the

eogenetic (Vacher and Mylroie 2002) Upper Floridan aquifer, which has approximately

30% primary porosity within the matrix rocks (Budd and Vacher, 2004). Madison Blue

Spring is a first magnitude spring (> 2.3 m3/sec; Meinzer, 1927), discharging up to 6.3

m3/sec at velocities up to 0.90 m/s from >7 km of mapped and additional unmapped

conduits. Peacock Springs is not a true spring with perennial discharge, but instead a

group of karst windows (including sampling sites Peacock 1 and Orange Grove) (Fig. 2-

1B) linked by mapped conduits. Mapped conduit lengths are 12.6 km and exploration is

continuing. At the closest approach, the cave is within two kilometers of the Suwannee

River.

Both springs receive inflow from matrix porosity during baseflow conditions and

neither spring is linked to sinking streams. Both springs have water with similar

chemical compositions during baseflow that reflects equilibration with calcite. River

water intrudes both springs as stage increases above the elevation of the potentiometric

surface at the spring vent. At the Madison Blue Spring, intrusions occur at various river

elevations depending on antecedent groundwater heads. At Peacock Springs, intrusion

of the primary karst window starts when the river stage reaches ~8 m above sea level

(masl) to top a sill separating the karst windows from the river. Which karst windows

receive river water from overland flow during flooding depends on the magnitude of the

Page 25: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

25

flood, with the Orange Grove site requiring larger flood events to receive water than

does the Peacock 1 site.

2.2.2 Methods

2.2.2.1 Field methods and sample analyses

Daily average river stages during the floods were obtained from two US

Geological Survey (USGS) gauging stations (http://waterwatch.usgs.gov/?m=real&r=fl).

These stations include Blue Spring Station (USGS Gauge Station 02319302) at

Madison Blue Spring and Luraville Bridge Station (USGS Gauge Station 02320000)

near Peacock Springs. These stations are the nearest river stage monitoring points to

the cave systems.

Sensors were placed in the rivers, springs, and conduits to record water

characteristics. In Situ Multi-Parameter Series TROLL 9500s were placed inside both

springs, about 30 m upstream of the spring vent. These sensors recorded specific

conductance and temperature at 30-minute intervals and were recalibrated

approximately monthly during the study. Schlumberger conductivity, temperature, and

depth (CTD) Diver loggers were placed in the Suwannee and Withlacoochee Rivers

upstream of Peacock and Madison Blue Spring systems, respectively, as well as at

various points in the cave systems. CTDs logged approximately every 30 minutes and

were calibrated when the data were downloaded, approximately every six months.

Variations in specific conductance and temperature in river and cave waters were used

to track the timing and extent of intrusion events.

Water samples were collected from rivers, spring vents, and karst windows

during four high-flow events and their subsequent recessions. These events raised

river elevations enough to intrude river water into the spring vents, but not all stages

Page 26: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

26

were sufficient to overtop their banks and thus were not true floods. For simplicity,

however, these events will be referred to as floods. Sampling typically began a few

days after the start of intrusion. At Peacock Springs, six samples were collected at the

Orange Grove karst window and from the Suwannee River at the Luraville gauging

station between 16 April and 14 July 2009 (reported in Gulley et al., 2011) and seven

samples were collected from the Peacock 1 karst window and from the Suwannee River

at the Luraville gauging station between 3 February and 1 April 2010. At Madison Blue

Spring, seven samples were collected between 6 April and 9 May 2011 and 17 samples

were collected between 5 March and 15 May 2012 from the spring vent and

Withlacoochee River. Each sampling event is referred to here by the year in which

sampling took place (e.g., 2009 flood, 2010 flood, etc.).

Samples were collected using a peristaltic pump connected to flexible PVC

tubing extended with rigid PVC pipe from the edge of the water body (spring vent, karst

window, or river). Water was pumped into a flow-through cell that held a sonde

connected to a calibrated YSI 556MPS instrument that measured pH, dissolved oxygen

(DO), and specific conductance (SpC). Water was pumped over the sonde at low flow

rates for at least eight minutes until all values stabilized. Filtered water samples were

collected in acid-washed, 20- mL plastic screw-top bottles and preserved with trace-

metal-grade nitric acid to pH < 2 for subsequent analyses of major cation concentrations

(Na+, K+, Mg2+, and Ca2+). Filtered water samples were collected unpreserved in new,

but not acid-washed, 20- mL plastic screw-top bottles to measure major anion

concentrations (Cl-, SO42-). Dissolved organic carbon (DOC) samples were collected in

pre-rinsed, 40- mL ashed amber glass bottles and preserved with HCl. Dissolved

Page 27: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

27

inorganic carbon (DIC) samples were collected in 20-mL French square glass bottles

and treated with three drops of saturated HgCl2 to prevent microbial activity. All samples

were kept on ice in the field and refrigerated after returning to the lab.

Dissolved inorganic carbon concentrations were measured by acidifying water

samples using an AutoMate Prep Device plumbed to a UIC (Coulometrics) 5011 carbon

coulometer, which measured the evolved CO2. The method was standardized with

known quantities of dissolved KHCO3 and generated data accuracy better than 1% on

all sample measurement runs. Dissolved organic carbon concentrations were measured

with a Shimadzu TOC-5000A total organic carbon analyzer by sparging samples for 2

minutes with C-free air to remove inorganic C. After high temperature combustion of

the organic carbon (OC), CO2 was measured through infrared analysis. The coefficient

of variance was <5% for replicate injections of each sample, and values reported here

are means of all DOC sample injections for individual samples. Cation and anion

concentrations were measured with an automated Dionex DX500 Ion Chromatograph.

Of the 74 samples collected in this study, 69 had a charge balance error < 10% and all

errors were positive. The five samples with charge balance errors > 10% (also positive)

were collected near the peak of floods, resulting in low ion concentrations, and had

DOC concentrations in excess of 10 mg/L. DOC concentrations exceeding 10 mg/L

contribute an unquantified negative charge to the system waters (Cantrell et al., 1990,

Hemond, 1990) affecting charge balance.

2.2.2.2 Estimates of mixing of river water and groundwater

The relative fractions of river water and groundwater in conduits were estimated

through each flood recession based on two-end-member mixing of Cl concentrations of

intruding river water and ground water (Brown et al., 2014; Brown et al. in prep). The Cl

Page 28: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

28

concentrations were estimated to be 0.25 mM ± 3% at Madison Blue and 0.20 mM ± 3%

at Peacock for the intruding river water end-member, and 0.16 mM ± 3% at Madison

Blue and 0.17 mM ±3 % at Peacock for the pre-intrusion water end-member. This model

assumes that Cl is conservative in this system, which is likely because no Cl-bearing

minerals are known from the region. Cl concentrations vary through time when spring

water is completely displaced by intruding river water in the conduits, limiting the

accuracy of mixing model estimates at the beginning of the recession (Brown et al.,

2014).

2.2.2.3 Dissolution rate estimates

We focused our geochemical modeling on carbonate dissolution through time

using two independent rate models and coefficients. One model was developed by

Plummer et al. (1978, herein referred to as the PWP model). The PWP model

estimates the rate, R, of the overall reaction mechanism, based on empirically derived

rates for each step in the carbonate dissolution reaction, according to:

𝑅 = 𝑘1𝑎𝐻+ + 𝑘2𝑎𝐻2𝐶𝑂3+ 𝑘3𝑎𝐻2𝑂 + 𝑘4𝑎𝐶𝑎2+𝑎𝐻𝐶𝑂3

− (2-3)

where k represents a temperature-dependent reaction rate and 𝑎 is the activity of

individual reaction species. Each value of k is calculated from dissolution

measurements of Icelandic Spar under conditions of pCO2 ranging from 0.0 to 1.0 atm

and temperatures between 5° and 60° C (Plummer et al., 1978). The first term is mass-

transfer controlled and thus depends on flow rate, but the second and third terms are

controlled by reaction rate. Therefore, the total dissolution rate is only partly controlled

by flow regime.

Page 29: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

29

The second model is modified from Svensson and Dreybrodt (1992; herein

referred to as SD model). This model uses ratios of measured Ca2+ concentrations and

estimates of Ca2+ concentrations of water that would be at equilibrium with calcite

(herein referred to as “equilibrium Ca2+ concentrations”) to estimate R according to:

𝑅 = 𝛼(1 − 𝐶/𝐶𝑠)𝑛 (2-4)

where C is the measured Ca2+ concentration and Cs is the equilibrium Ca2+

concentration, n is the empirical reaction order (unitless), α is the laboratory-derived

dissolution rate (mmol cm−2 s−1). The equilibrium concentration of the water with

respect to calcite was calculated using the phreeqc.dat database in PHREEQc

(Parkhurst and Appelo, 1999) and all measured solute concentrations. The α values

from Svensson and Dreybrodt (1992) were derived from dissolution experiments

conducted in an environment at 20° C with pCO2 at 5 x 10-3 atm, and include a range of

Ca2+ concentrations that gives C < xCs, with x ≈ 0.8. Both α and n values vary with the

saturation state of the water. When x > 0.6, the empirical reaction order (n) is 3–4

(Svensson and Dreybrodt, 1992). For values of x < 0.6, n ≈ 1.5−2.3. For x > 0.3, α ≈ 1.6

to 2.2 x 10−7 mmol cm−2 s−1. For x < 0.3, α ≈ 3 x 10−6 (Svensson and Dreybrodt, 1992;

Kaufmann and Dreybrodt, 2007). The n and α values used in this study were the

averages of the range of values when not explicitly stated. Averaging the range of

values appears appropriate based on a sensitivity analysis that indicates variations in n

and α varied rates by ~70% over the full range of compositions of flood waters, but only

by around 30% when most dissolution occurs near the start of the floods.

Model equations 2-3 and 2-4 were used to estimate dissolution rates for both

flood and baseflow samples with units of mass per area per unit time (mmol/cm2/day).

Page 30: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

30

The amount of dissolution for each sample was estimated by multiplying that sample

dissolution rate by the time interval between sampling events with a right Riemann sum

extending to the date of the measurement. Dissolution for each flood was calculated by

summing dissolution amounts from the initial sample to the return to baseflow. The

flood dissolution amounts are considered minima because of the lack of samples during

the first few days of each flood and the use of a right Riemann sum, which should

slightly underestimate the amounts.

2.2.2.4 Impacts of mixing/dissolution Ca2+ additions and DOC remineralization on carbonate dissolution

Dissolution rates in natural systems slow when Ca2+ concentrations increase as

calcite dissolves and Ca2+-rich water mixes with Ca2+-poor water. We examined

whether dissolution or mixing slows dissolution more by using two data sets: (1)

measured sample ion concentrations and (2) ion concentrations estimated from

conservative mixing models. The first is comprised of measured ion concentrations of

samples collected during flood recession described in section 2.2.1. The mixed

composition of waters is determined from a weighted average of fractions of flooded

river water and pre-intrusion groundwater ion concentrations described in section 2.2.2.

These two data sets will be referred to as “measured model” and “mixed model” in this

paper. The models are calculated only for the 2010 and 2012 events because these

events experienced complete displacement of conduit waters, represent both sample

locations, and were sampled at sufficiently high resolution. The measured model

represents Ca2+ concentrations that result from both mixing and in situ calcite

dissolution, whereas the mixed model represents Ca2+ concentrations that result solely

from mixing, excluding in situ calcite dissolution. We assume the differences between

Page 31: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

31

the measured and mixed models represent the magnitude of in situ dissolution.

Because both the measured and mixed model show water remained undersaturated

during the recession, the remaining potential dissolution capacity of the water to

dissolve calcite was estimated by using PHREEQc to react the measured model and

mixed model water compositions to a SIcal = 0. These solutions will be referred to as

“measured SI 0” and “mixed SI 0.” If the mixed model predicts more potential dissolution

than the measured model, we assume the extra potential dissolution results from in situ

remineralization of OC.

2.3 Results

Temperature and water compositions, as reflected in the SpC, changed

systematically throughout each intrusion event regardless of its size, duration, or

season (Fig. 2-2). Generally, SpC decreased more in large than small floods and

temperature decreased more during winter than summer floods. These changes in SpC

and T correspond with changes in the Ca2+ concentrations and SIcal values. The SIcal

values were low at the start of intrusion and decreased to vanishingly small values (6 x

10-18 mmol/cm2/day) at baseflow (Fig. 2-3). The return to near-equilibrium conditions

corresponds to slow dissolution rates (Fig. 2-4). Although the dissolution patterns are

generally the same in each event, the magnitudes of the floods and hydrologic

characteristics differ at each spring, which affect the dissolution rates and total calcite

dissolved. These differences are described below for each spring.

2.3.1 Peacock Springs

During the 2009 flood, the river stage increased by more than 8 m,

simultaneously with decreases in SpC of ~365 µS/cm (409 to 44 µS/cm) and

temperature of ~4° C (21.7 to 17.7° C) (Fig. 2-2). The 2010 flood was smaller, with an

Page 32: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

32

increase in stage of 3.5 m and decreases in SpC of ~ 270 µS/cm (322 to 53 µS/cm) and

temperature of ~12° C (20.7 to 8.4° C). This flood occurred in winter, causing the large

decrease in temperature. Chloride mixing models and temperature data indicate water

in the conduits was completely displaced by intruding river water during both floods.

The 2009 flood was characterized by a minimum SIcal value of -6.43, whereas the

2010 flood had a minimum SIcal value of -5.27 (Fig. 2-3 and Table 2-2). The SIcal values

during the 2009 flood reflect dissolution rates that ranged from 0.15 to 0.20

mmol/cm2/day and total dissolution of 1.60 and 2.01 mmol/cm2 based on the SD and

PWP models, respectively (Table 2-2). Water in this event had SIcals of -0.5 and -1 for

most of the flood recession (Fig. 2-6A). During the 2010 flood, the maximum dissolution

rates were calculated to be 0.007 and 0.02 mmol/cm2/day (Fig. 2-4) with total

dissolutions of 0.10 mmol/cm2 and 0.75 mmol/cm2 (Fig. 2-5 and Table 2-2), as

estimated by the SD and PWP models, respectively. This flood had SIcal < -1 for more of

the recession than any other flood (Fig. 2-6B).

Both the mixed and measured models predict that Ca2+ concentrations increase

with time following both intrusions (Fig. 2-7). Following the 2010 flooding, measured

Ca2+ concentrations were ~59% greater than the mixed model estimated concentrations

(Fig. 2-7C). Measured Ca2+ concentrations gained ~11% more Ca2+ when equilibrated

to SI 0. Mixed Ca2+ concentrations gained ~64% more Ca2+ after being equilibrated

with calcite (Fig. 2-7A). During the 2010 flooding, the measured SI 0 data had ~ 8%

more Ca2+ than mixed SI 0 (Fig. 2-7E).

2.3.2 Madison Blue Spring

During the 2011 flood, the river stage increased by ~1 m simultaneously with

decreases in SpC of ~45 µS/cm (from 304 to 257 µS/cm) and temperature of ~0.5° C

Page 33: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

33

(20.9 to 20.4° C) (Fig. 2-2). These small changes in both SpC and temperature values

indicate that only 26% of the conduit spring water was displaced by intruding river

water. The 2012 flood was larger, with an increase in stage of ~2.5 m, simultaneous

with decreases in SpC of ~250 µS/cm (from 345 to 86 µS/cm) and temperature of ~5.1°

C (20.6 to 15.5° C), indicating all conduit water was displaced (Fig. 2-2D).

The 2011 flood was characterized by a minimum SIcal value of only -1.19,

whereas the SIcal value decreased to a minimum of -4.73 in the 2012 flood (Fig. 2-3 and

Table 2-2). These differences in saturation states are reflected in low dissolution rates

during the 2011 flood of 0.0004 to 0.01mmol/cm2/day as calculated by the SD model

and PWP models, respectively (Table 2-2). Dissolution during floods ranged from

0.0004 to 0.37 mmol/cm2 based on the SD and PWP models, respectively. During the

entire flood, SIcal ranged between 0 and -1, with values between SIcals of 0 and -0.5 for

~95% of that time. During the 2012 flood, maximum dissolution rates at Madison Blue

Spring were 0.14 (Fig. 2-4) to 0.02 mmol/cm2 based on the SD and PWP models,

respectively. Dissolution during the flood ranged from 0.60 to 0.81 mmol/cm2 based on

the SD and PWP models, respectively. This flood had SIcal values between 0 and -1 for

approximately 90% of the time, and values varied between SIcals of -1 to -4 the rest of

the time (Fig. 2-6B).

Concentrations of Ca2+ increased with time following spring reversals for both the

mixing model and measured data sets (Fig. 2-7). Following the 2012 flooding,

measured Ca2+ concentrations were ~8% greater than mixed concentrations (Fig. 2-

7D), but gained ~8% when equilibrated to SI 0 (Fig. 2-7B). Mixed Ca2+ concentrations

gained ~ 12% more Ca2+ after being equilibrated with calcite (Fig. 2-7B). During the

Page 34: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

34

2012 flooding, the measured SI 0 data had ~ 4% more Ca2+ than the mixed SI 0 (Fig. 2-

7F).

2.4 Discussion

The amount of calcite that dissolves as surface water intrudes karst aquifers

depends on the volume of intruding water, the residence time in the aquifer of intruded

water, its undersaturation with respect to calcite, and dissolution reaction rates. Both

the volume of intruding river water and the residence time in the aquifer depend on the

magnitude of difference in hydraulic head gradients and the length of time they are

oriented from the river to the aquifer (Gulley et al., 2011). This return to nearly

equilibrated water could result from mixing with near-calcite-saturated and low-DOC

groundwater, dissolution of calcite by the intruding water during the intrusion, less

production of CO2 by remineralization of DOC as energetically favored terminal electron

acceptors (e.g., O2) are consumed, or a combination of these factors. For the following

discussion, we used water compositions and changes in saturation state to estimate

reaction rates and magnitudes of dissolution during the flood.

2.4.1 Temporal Variations in Dissolution Rates

The four sampled intrusion events exhibited different maximum dissolution rates,

total dissolution (Table 2-2) and temporal variations in rate. Most dissolution occurs

early in the floods as dissolution rates decrease during recessions; more than 90% of

calcite dissolution occurred within the first 20% of all floods (Fig. 2-4). The return to

slow reaction rates during flood recession reflects water saturation states being near

equilibrium with calcite at the approach to baseflow. Variations in maximum rates result

from differences in the flood magnitudes, which correspond to the maximum amount of

undersaturation (Fig. 2-3). In addition to lower saturation states, large floods force more

Page 35: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

35

river water into the spring system over a longer period of time than do smaller floods.

Large intrusion events also force more water into the rock matrix from the conduits

(Martin and Dean, 2001; Moore et al., 2010), thereby increasing the surface area with

which undersaturated water reacts.

2.4.2 Comparison of Dissolution Rate Models

Dissolution rates and magnitudes vary between floods, but within individual

floods, the rates depend on the model used to calculate them. The SD model (equation

2-4) estimated less dissolution than the PWP model (equation 2-3), except for

Suwannee River during the 2012 flood (Fig. 2-5). The SD model estimates are ~80% of

the PWP model estimates during the 2009 flood, ~14% of the PWP model estimates

during the 2010 flood, and ~0.12% of the PWP model estimates during the 2011 flood.

These differences result from differences in model behavior near equilibrium and at

different pCO2 values. The PWP model is appropriate for water that is far from

saturation with respect to calcite (<20% of the Ca2+ concentration when at SIcal = 0)

(Plummer et al., 1979; Buhmann and Dreybrodt, 1985), but rates may be overestimated

in waters near SI 0 with high ion concentrations (Svensson and Dreybrodt, 1992), which

may inhibit dissolution by retarding acids from reaching reaction sites. Consequently,

the SD model may better characterize dissolution than the PWP model in waters close

to equilibrium with calcite, which is the normal for most of the recession following the

intrusion and rapid shift toward equilibrium with respect to calcite (Fig. 2-4). The SD

model was developed using 5 x 10-3 atm of CO2 and the samples we measure have

pCO2 values that ranged between 0.0009 and 0.15 atm. The changing pCO2 should

also alter the Cs value in the SD model, therefore accounting, at least partially, for

differences in the modeled and measured samples. Because the PWP coefficients

Page 36: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

36

were determined at pCO2 pressures up to 1 atm, results from this model may be more

accurate at the high pCO2 concentrations seen during some portions of the study.

The PWP model uses several different temperature-dependent rate constants

and the first term is mass-transfer controlled, whereas the second and third terms are

controlled by the reaction rate. This equation design accounts for both differences in

flow and reaction rates. In contrast, the SD model relies solely on the calcite saturation

state of water to determine reaction rates, and the exponent in the SD model (eq. 2-4)

creates abrupt changes in rates as saturation thresholds are crossed. Each model thus

predicts different amounts of total dissolution depending on the extent of

undersaturation experienced and length of time spent at different levels of

undersaturation (Fig. 2-6 B). In the case of a flood that does not result in highly

undersaturated conduit waters (e.g., the 2011 flood), the PWP model estimates more

dissolution than the SD model because of its assumption of linearity between

dissolution and saturation state even when water is near equilibrium with calcite. In

contrast, the SD model estimates reaction rates ~4 orders of magnitude slower than the

PWP model when water is between SIcal 0 and -1 (Fig.2- 6 A).

2.4.3 Saturation States and Dissolution Magnitude Through the Recession

The greatest undersaturation occurs at the beginning of intrusion events because

of CO2 produced by OC remineralization and lack of water contact with carbonates in

the river channel. Ground water near equilibrium with calcite mixes with the intruded

river water as hydraulic head gradient differences decrease and intrusion slows. Mixing

slows dissolution rates because the ground water is near equilibrium, but in situ

dissolution also diminishes the undersaturation and slows dissolution rates. The

following discussion estimates the relative importance of mixing, in situ dissolution, and

Page 37: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

37

in situ organic carbon remineralization in controlling dissolution rates and magnitudes of

dissolution.

Both the mixed and measured models show trends toward equilibrium during the

2010 and 2012 floods (Fig. 2-7 A and B). The measured model Ca2+ concentrations are

always greater than mixed model Ca2+ concentrations, indicating that in situ dissolution

has released Ca2+ in excess of that derived from mixing. Subtracting the mixed Ca2+

concentrations from the measured Ca2+ concentrations yields the Ca2+ contributed from

dissolution This Ca2+ is ~59% and 8% of the mixed Ca2+ concentrations during the

2010 and 2012 floods, respectively. The greater Ca2+ contributed in 2010 and 2012

likely derives from the length of intrusion and differences in the hydrology of the springs.

Intruded river water stopped draining from Peacock Springs after the water table

dropped below 8 m above sea level and thus remained in the aquifer longer than at

Madison Blue Spring.

Because the mixed water samples contain less Ca2+ than the measured water

samples, they gain more total Ca2+ when equilibrated with calcite to SIcal = 0 than the

measured water samples. Even when equilibrating to SIcal = 0, the total Ca2+

concentration of the mixed SI 0 remains less than the measured SI 0. The difference

between these two estimates represents chemical reactions other than carbonate

dissolution that affect ability of water to dissolve calcite, most likely from the

remineralization of DOC. Subtracting the value of the mixed SI 0 from the measured SI

0 indicates that DOC remineralization adds ~8% and ~4% more Ca2+ in the measured

model SI 0 than the mixed model SI 0 during the Peacock 2010 flood and Madison Blue

2012 flood. The addition of CO2 through DOC remineralization continues to drive

Page 38: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

38

undersaturation until the DOC is gone or the remaining fraction is completely

recalcitrant. This prolonged addition of CO2 may allow further dissolution in areas

where water remains trapped longer, such as Peacock Springs, than in areas that

discharge intruded waters quickly, such as Madison Blue Springs. The sill that prevents

Peacock Springs from reversing as frequently as Madison Blue Springs may have

contributed, in part, to Peacock Springs reaching a greater length than Madison Blue by

retaining river water longer in the system, thereby driving greater dissolution.

2.5 Conclusions

This study presents the first estimates of dissolution rates in karst aquifers

caused by changes in water chemistry during intrusions caused by flooding. Estimates

of rates based on two models with distinct assumptions and with data collected during

four floods indicate dissolution rates increase by up to 17 orders of magnitude during

floods compared to rates at baseflow, indicating that most dissolution in these systems

is episodic. The two models yielded different estimates of dissolution during floods.

Highest dissolution rates occurred during the first 20% of the flood, when more than

80% of the dissolution occurred. Dissolution rates are elevated at the start of intrusions

because of the undersaturated state of the intruding river water with respect to calcite,

which is enhanced by remineralization of DOC that intrudes with the floodwater.

Dissolution rates slow near the end of intrusions as the intruded river water begins to

discharge. Diffuse recharge of rainfall through the land surface, although also

undersaturated with respect to calcite, causes little dissolution of the conduits, which

expand episodically during intrusion events. At the beginning of a flood, dissolution is

almost the exclusive source of Ca2+ added to intruding river water, but when the flood

Page 39: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

39

recession begins, simple mixing adds Ca2+ to conduit waters and is primarily

responsible for slowing dissolution rates.

Page 40: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

40

Table 2-1. Karst dissolution rates.

Analysis Type Method Rate (mm/yr) Authors

Theoretical (Laboratory

Based)

Peak Rates using linear Dissolution Rate

Equations (varying CO2 concentrations)

2.3 – 16.3a Plummer et al., 1978

Peak Rates from Dissolution Rate

Equations similar to this study (open system varying water film

thickness)

0.09 - 0.5 Buhmann and

Dreybrodt, 1985 a

Peak Rates from Dissolution Rate

Equations similar to this study (closed system

varying water film thickness)

0.3 -1.45 Buhmann and

Dreybrodt, 1985 b

Atomic Force Microscopy 8.54 x 10-8 - 8.54

x 10-4 Dove and Platt 1996

Atomic Force Microscopy 4.2 x 10-4 - 1.6 x

10-3 Shiraki et al., 1999

Observational (Field Based)

Measuring Post Glaciation Pedestals

0.015-0.040 Jennings 1985

Measured Ca2+ Concentrations in

Mountain Drainage Basins 0.075-0.083 Pitty 1968

Estimated Ca2+ Concentrations in Springs

0.03 Opdyke et al., 1984

Geomorphic estimations of denudation

0.09 Adams et al., 2010

Baseflow Conditions 1 x 10-16 This study

Flood Conditions 0.000019 – 0.08b This study aRates are taken from the greatest potential values from linear rate calculations and may not be representative of most natural systems. bRates are based solely on the SD equation and do not account for DOC remineralization (see text for explanation).

Page 41: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

41

Table 2-2. DOC concentrations, SI values, and dissolution rates for the four sampled floods.

Location Year

Peak Intruding DOC mg C/L

Lowest SIcal Value

SD Maximum Dissolution Rate (mmol/cm2/day)

SD Total Dissolution during Flood (mmol/cm2)

PWP Maximum Dissolution Rate (mmol/cm2/day)

PWP Total Dissolution during Flood (mmol/cm2)

Peacock 2009 28.39 -6.43 0.15 1.60 0.20 2.01

Peacock 2010 24.3 -5.27 0.007 0.10 0.02 0.75

Madison 2011 2.07 -1.19 4.1x10-4 4.3X10-4 0.01 0.37

Madison 2012 19.11 -4.73 0.14 0.60 0.02 0.81

Page 42: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

42

Figure 2-1. Maps of river and cave systems. A) Map of the Suwannee River Basin showing locations of Peacock and Madison Blue Springs cave systems and stage gauges on the Suwannee River. B) Map of Peacock Springs cave system. C) Map of Madison Blue Spring cave system in relation to the Withlacoochee River. Modified from Gulley et al. (2011).

Page 43: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

43

Figure 2-2. Conductivity and temperature through time at spring vents for reversals: A) Peacock 2009, B) Peacock 2010, C) Madison Blue 2011, D) Madison Blue 2012. The y axes are fixed at the same range, but have different values, to emphasize differences in magnitudes between the events. Arrows at the top of graphs indicate sampling times. The sharp drop in temperature and specific conductivity marks the initial intrusion of river water into the system followed by a recession to base flow.

Page 44: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

44

Figure 2-3. Concentrations of calcium and calcite saturation indices through time for A) Peacock 2009, B) Peacock 2010, C) Madison Blue 2011, and D) Madison Blue 2012. Concentrations of calcium are laboratory measured values of grab samples and saturation indices were calculated in PHREEQc.

Page 45: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

45

Figure 2-4. Time variations in the river stage, dissolution rates for reversals, and cumulative dissolution at A) Peacock 2009, B) Peacock 2010, C) Madison Blue 2011, D) Madison Blue 2012. Estimated dissolution rates are shown for the spring vent (solid points) and the river adjacent to the springs (open points). Plotted rates are estimated based on the SD model only (see text for discussion). Cumulative dissolution is shown as a dashed black line. River stages are daily averages shown as a solid grey line. Sample points indicate dissolution rates at that time; dissolution likely started before the first samples were collected. Vertical black lines through a point indicate that the ρCO2 of the sample was larger than 0.005 atm (the atm at which the SD equation was tested). Cumulative dissolution incrementally sums the area under the curve plotted for daily dissolution rates through an event.

Page 46: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

46

Figure 2-5. Total calcite dissolution as measured by Ca2+ for each event by the SD equation (black) and the PWP equation (Gray). The river value represents values at Withlacoochee River for Madison Blue Spring and Suwannee River for Peacock Spring. The letter P represents Peacock and MB represents Madison Blue Spring.

Page 47: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

47

Figure 2-6. Predicted calcite dissolution values and SI range durations. A) Dissolution rates using the PWP and SD equations at a range of saturation indices with respect to calcite. B) Cumulative fraction of time that each flood spent in a given range of saturation indices with respect to calcite. Ranges cover saturation index 0 to -4 in 0.5 saturation index units each and all samples under -4 are labeled as “under 4.”

Page 48: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

48

Figure 2-7. Measured and mixed Ca2+ concentrations and measured SI 0 and mixed SI 0 Ca2+ concentrations through time. A) Peacock 2010 and B) Madison Blue 2012. Estimates of Ca2+ added by carbonate dissolution from measured Ca2+ concentrations minus mixed Ca2+ concentrations through time at C) Peacock 2010 and D) Madison Blue 2012. Estimates of potential Ca2+ additions from organic carbon remineralization based on measured SI 0 Ca2+ concentrations minus mixed SI 0 concentrations through time at E) Peacock 2010 and F) Madison Blue 2012.

Page 49: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

49

CHAPTER 3 RELATIVE AMOUNTS OF DISSOLUTION IN CARBONATE TERRAINS FROM RIVER

LOSSES AND DIRECT RAINFALL

3.1 Introduction

Landscape evolution is determined by uplift, physical and chemical weathering,

bedrock lithology, climate, soil type and thickness, and vegetation (Ford et al., 1988;

Hoorn et al, 2010; Huggett, 2007; Mann, 2002). Landscape evolution tends to be more

rapid in carbonate than ice-free siliciclastic landscapes because dissolution, rather than

physical erosion, provides the major control (Huggett, 2007). Carbonate dissolution

forms karst landscapes, which are characterized by caves, sinking streams, and dolines

(White, 1988; Ford and Williams, 2007). This chemical erosion is especially prevalent in

middle to low latitudes where rainfall and vegetation are plentiful. Dissolution is

primarily driven by carbonic acid, though other naturally occurring acids, such as sulfuric

acid, can also cause dissolution (Botrell et al., 1991; Hercod et al., 1998; Johnson and

Hallberg, 2005; Spence and Telmer, 2005). Carbonic acid forms as CO2 in the

atmosphere and soils dissolves in rainwater,

𝐶𝑂2 + 𝐻2𝑂 ↔ 𝐻2𝐶𝑂3 (3-1)

which then dissolves soluble bedrock. Where calcite dissolves, calcium and

bicarbonate concentrations are elevated in surface water and groundwater

𝐻2𝐶𝑂3 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 2𝐻𝐶𝑂3− (3-2)

These elevated ion concentrations are then transported through the hydrologic systems.

The hydrologic system is important in determining where dissolution occurs by

controlling the locations of contact between limestone and carbonic acid (Fig. 3-1).

Page 50: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

50

Hydrology also controls the magnitude of dissolution by altering flow rates, residence

times and the extent that reaction products are flushed from the aquifer. Flow through

the subsurface may concentrate dissolution in areas to form caves, referred to as

conduits when water filled. Concentrated dissolution occurs at preferential flow paths

because of water piracy/flow capture (Szymczak and Ladd 2011) and at aquitards and

geologic structures, such as bedding planes or joints (Palmer, 1991). Concentrated

dissolution also occurs at the water table (Florea et al. 2007, Gulley et al. 2013). As

rivers flood, undersaturated water may penetrate river banks and enter the aquifer and

cause dissolution (Chen and Chen, 2003). Such penetrating flow can occur at large

spatial scales in karst systems when elevated river hydraulic heads reverse spring flow,

forcing large volumes of undersaturated water into the aquifer (Gèze, 1987; Albéric and

Lepiller, 1998; Gulley et al., 2011). During these events, limestone dissolution rates in

phreatic conduits can be up to 16 orders of magnitude faster than rates under baseflow

(Fig. 2-4), and these reactions are spatially constrained by the penetration depth of river

water into the rock matrix. Dissolution may also be distributed regionally across the

land surface as rainfall (Fig. 3-1C), which derives CO2 from the atmosphere and soils, is

undersaturated with respect to carbonate minerals (eq. 3-2).

Estimates of the magnitude of dissolution can be converted to rates at which the

land surface is lowered (i.e., denudation rates) by normalizing the dissolution

magnitudes to defined areas. Estimates of denudation use a variety of methods

including measuring armored pedestal heights above the surrounding land (Jennings,

1985), calculating dissolution required to produce measured spring and river Ca2+

concentrations (Pitty, 1968; Opdyke et al., 1984) and calculating dissolution rates

Page 51: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

51

necessary to yield uplift (Adams et al., 2010). Dissolution-driven denudation was

compared with stream erosion and found to occur at approximately the same rate in the

tectonically active Appalachian karst region (White, 2009). Similar to Appalachian karst

(Granger et al., 2001), caves occur at several different depths below the land surface in

the Florida karst (Florea et al., 2007). Dissolution in north-central Florida is

hypothesized to have caused isostatic uplift equal to ~36 m during the Pleistocene and

Holocene, resulting in denudation rates of between 0.03 and 0.09 mm/yr (Opdyke et al.

1984; Adams et al., 2010). In some areas, rivers have down cut sufficiently to bisect

phreatic caves that were previously unconnected to surface flow (Gulley et al. 2013).

What remains unknown is the relative magnitude of regional denudation from rainfall

across the land surface versus concentrated subsurface dissolution, processes that

gave rise to the modern Florida landscape.

We examined one rainfall event (Tropical Storm Debby) that passed through

north-central Florida in 2012 and was sufficient to cause flooding and trigger a spring

reversal. We estimated the amount of dissolution caused by rainfall recharge across

the land surface, along with recharge volumes from the spring intrusion and bank

penetration, which allowed us to estimate the relative magnitudes of these three

dissolution mechanisms for the first time. By comparing these magnitudes, we were

able to assess how these processes have shaped the hydrology and geomorphology of

the region.

3.2 Study Sites and Methods

3.2.1 Study Sites

Located in north Florida, Madison Blue Spring discharges to the Withlacoochee

River ~17 km downstream of the Pinetta gauging station and ~ 15 km upstream of the

Page 52: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

52

Lee gauging station (Fig. 3-2). The region receives an average of 1334 mm of annual

rainfall (NCDC 2015), with approximately half falling in summer during localized

thunderstorms and tropical storms. Madison Blue Spring is located near the Cody

Scarp, a regional geomorphic feature that marks the erosional edge of the Miocene

Hawthorn Group (Fig. 3-2A), which confines the Upper Floridan aquifer. The river reach

from the Pinetta gauging station to Madison Blue Spring flows across the Hawthorn

Group, where it is sufficiently thin to allow water to be lost to the Upper Floridan aquifer

from the river. The river reach from Madison Blue Spring to the Lee gauging station

flows across the unconfined Upper Floridan aquifer and is a gaining river. The

Withlacoochee River is tannic and sourced primarily from wetlands, agricultural fields,

and forested land runoff.

Madison Blue Spring discharges from >7 km of mapped conduit system;

exploration is on-going (Fig. 3-2B). Madison Blue Spring is a first magnitude spring (>

2.3 m3/sec; Meinzer, 1927), discharging up to 6.3 m3/sec at velocities up to 0.90 m/s

(SJRWMD). Conduits that source Madison Blue Spring occur in the eogenetic (Vacher

and Mylroie, 2002) Upper Floridan aquifer, which has approximately 30% primary

porosity within the matrix rocks (Budd and Vacher, 2004). Madison Blue Spring

contributes about 30% of the Withlacoochee discharge (Farrell and Upchurch 2004).

Water at Madison Blue Spring reverses flow when hydraulic head gradients drive flow

toward the spring, as river stage rises more rapidly than the groundwater hydraulic head

(Brown et al., 2015), which is a function of antecedent groundwater levels, recharge

from infiltration, and lateral flow.

Page 53: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

53

3.2.2 Methods

3.2.2.1 Monitoring data and historical reversal estimates

Several sets of monitoring data were used, including discharge of the

Withlacoochee River, rainfall amounts, and the chemical composition of rainfall. Daily

average river stage and discharge during the flood were obtained from three US

Geological Survey (USGS) gauging stations along the Withlacoochee River, including

Pinetta (USGS 02319000), near Madison (USGS 02319300), and Lee (USGS

02319394). Daily rainfall totals were downloaded from the Live Oak weather station

(Location 170), which is part of the Florida Automated Weather Network (FAWN)

(http://fawn.ifas.ufl.edu/) and located ~40 km ESE of Madison Blue Spring. Daily

average chemical composition of rainfall was downloaded from the Quincy weather

station (NTN site: FL 14), which is part of the National Atmospheric Deposition Program

(http://nadp.sws.uiuc.edu/data/NTN/) and located ~130 km west of Madison Blue

Spring.

Historic stage gauge data at Pinetta were used to identify when spring reversals

may have occurred at Madison Blue Spring over an ~80-year period of record. The

Pinetta gauging station was selected because it was near the Madison Blue spring

system, and the gauging station at Madison Blue spring run only has records for the last

~14 years. Stage data were evaluated to find periods when characteristics were similar

to those during four known reversals (Table 2-2), including the one described here.

These characteristics include stage increases at a rate > 0.45 m/d for three consecutive

days, reaching a stage in excess of any stage in the previous two weeks, and a stage at

least 150% greater than the average stage at base flow during the previous year.

Baseflow was established for the year preceding the events by averaging low flow

Page 54: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

54

stage. One year should represent antecedent baseflow conditions because springs in

the area have been shown to respond to extreme weather events (e.g., tropical storms)

in approximately 6-8 months (White 2005; Florea and Vacher 2006).

3.2.2.2 Field methods and sample analyses

An In Situ Multi-Parameter Series TROLL 9500recorded specific conductance

and temperature at 30-minute intervals approximately 30 meters upstream of the

Madison Blue Spring vent. This logger was removed, maintained, recalibrated, and

immediately replaced approximately monthly during the study. Schlumberger

conductivity, temperature, and depth (CTD) diver loggers were placed in the

Withlacoochee River upstream of Madison Blue Spring as well as at various points in

the cave system. The CTDs logged approximately every 30 minutes and were

calibrated when the data were downloaded, approximately every six months.

Water samples were collected from the spring vent and Withlacoochee River 17

times between 5 March and 15 May 2012 using a peristaltic pump connected to flexible

PVC tubing extended with rigid PVC pipe from the edge of the water body (spring vent,

karst window, or river). Water was pumped into a flow-through cell that held a sonde

connected to a calibrated YSI 556MPS instrument that measured pH, dissolved oxygen

(DO), and specific conductance (SpC). Water was pumped over the sonde for at least

eight minutes until all values stabilized. Water samples were filtered through a 0.45-µm

trace-metal-grade canister filter and collected in acid-washed 20- mL plastic screw-top

bottles and preserved with trace-metal-grade nitric acid to pH < 2 for major cation

concentrations (Na+, K+, Mg2+, and Ca2+). Samples for major anion concentrations (Cl-,

SO42-) were collected unpreserved in new, but not acid-washed, 20- mL plastic screw-

top bottles. DOC samples were collected in pre-rinsed, 40- mL ashed amber glass

Page 55: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

55

bottles and preserved with HCl. DIC samples were collected in 20- mL French square

glass bottles and treated with three drops of saturated HgCl2 to prevent microbial

activity. All samples were kept on ice in the field and refrigerated after returning to the

lab.

Cation and anion concentrations were measured with an automated Dionex

DX500 Ion Chromatograph. Of the 34 sets of ion samples collected in this study, 33 had

a charge balance error < 10%, 29 had a charge balance error <5%, and all but three

sample sets had errors that were positive. The five samples with charge balance errors

> 5% were collected near the peak of floods, resulting in low ion concentrations, and

had DOC concentrations in excess of 10 mg/L. Dissolved organic carbon

concentrations exceeding 10 mg/L contribute an unquantified negative charge to the

system waters (Cantrell et al., 1990; Hemond, 1990) which results in charge balance

errors increasing and becoming more positive. Dissolved inorganic carbon

concentrations were measured by acidifying water samples using an AutoMate Prep

Device plumbed to a UIC (Coulometrics) 5011 carbon coulometer, which measured

evolved CO2. The method was standardized with known quantities of dissolved KHCO3

and generated data accuracy better than 1% on all runs. Dissolved organic carbon

concentrations were measured with a Shimadzu TOC-5000A total organic carbon

analyzer by sparging samples for 2 minutes with C-free air to remove inorganic C. After

high-temperature combustion of the organic carbon, CO2 was measured by infrared

analysis. The coefficient of variance was <5% for replicate injections of each sample,

and values reported here are means of all DOC sample injections for individual

Page 56: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

56

samples. All analyses were completed in the Department of Geological Sciences at the

University of Florida.

3.2.2.3 Recharge and dissolution modeling

We estimated dissolution potential in the three sub-regions where dissolution

may occur during flood events (Fig. 3-1). These include dissolution as water intrudes

conduits, dissolution longitudinally in river banks, and dissolution across the drainage

basin driven by rainfall which diffusely recharges the aquifer. These dissolution

estimates were made by first estimating the volume of water that enters each of these

sub-regions and then using water chemistry data to estimate the amount of dissolution.

Estimates of dissolution from each sub-region were compared to evaluate which area

contributes the most to dissolution overall, thereby altering and developing the karst

landscape and its hydrogeologic characteristics.

Three different approaches were used to estimate the volume of water flowing

into Madison Blue Spring during the 2012 reversal. 1) The first approach used the

Madison Blue Spring gauging station (USGS 02319302), which is located in the spring

run and recorded negative flow during the flood. 2) The second method used a

hydrologic model (Spellman, 2012) that utilized shifts in the hydraulic heads of the cave

system, wells, and the river to determine the volume of river water intrusion during the

spring reversal. The model simulated flow based on hydraulic gradients and estimated

hydraulic conductivities in the cave system and surrounding rock matrix by using the

finite difference model software package MODFLOW with the CFP extension. 3) The

third approach used a geochemical mixing model (Brown et al., 2015), which employs

Cl- concentrations and assumes conservative two-end-member mixing to calculate the

fractions of river water and groundwater contained in the post-reversal discharge. The

Page 57: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

57

river water fraction was multiplied by the total discharge volume based on discharge

records at Madison Blue gauging station to determine the river water discharge volume.

We recognize that the three methods employed to quantify volume are estimates and

rely on comparing these independently derived estimates with one another to find the

closest approximation of intruded river volume. We estimated the total dissolution that

occurred during the 2012 reversal using the Cl- mixing model to estimate the Ca2+

concentrations expected in discharge water that would result from simple mixing.

Measured concentrations were always in excess of the modeled concentrations and the

difference between the two is assumed to represent Ca2+ derived from calcite

dissolution. By multiplying the excess Ca2+ concentrations with the river intrusion

volumes, we calculated the total dissolution in Madison Blue Spring system during the

2012 spring reversal. These estimates are based on ion concentrations from grab

samples collected through time and therefore are subject to the ~5% error associated

some of these samples.

After the dissolution occurring during the spring reversal was calculated, we

calculated the intruded volumes necessary to determine how much dissolution occurred

along the studied river reach. To determine channel storage changes, LIDAR data

(SRWMD) were used to estimate the average channel volume per m around each

gauging station at various river stages. A stretch 1680 m long was used for averaging.

These entries created an index table with stages and volumes for the river reach.

Linear interpolation of the values in this table was used, in conjunction with daily stage

data, to estimate the channel water volume for the river reach each day. The volume

per meter from one day was subtracted from the volume per meter for the next day to

Page 58: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

58

establish the normalized change in volume per day. These estimates are limited by

small scale changes in channel geomorphology which will raise or lower channel

capacity at a given stage. While these variations are likely small, they do introduce

error into channel storage estimates.

This change in volume per meter was used as a “channel storage correction” to

account for daily water storage changes between the upper and lower gauging stations.

The gauging station discharge at the downstream gauging station was subtracted from

the upstream gauging station. The “channel storage correction” was subtracted

(assuming losses from the river to be positive) from the difference in gauge station

discharges to calculate the gain or loss of river water to the aquifer in the river reach per

day. The daily values were plotted and the area under the resulting curves was

computed to determine the volume of river water entering the aquifer in each reach as a

result of the storm. The volumes of river water that entered the aquifer (L) were

multiplied by the dissolution estimate used for the spring reversal (mmol/L) to calculate

the dissolution associated with bank penetration. River water loss estimates are

dependent on discharge calculations at the individual gauging stations; discharge is

based on a rating curve utilizing river stage and may actually vary due to changes in

channel geomorphology during the storm, which will introduce error into the

calculations.

Dissolution resulting from rainfall across the land surface was evaluated for the

drainage basin by estimating soil moisture from evapotranspiration (ET) and rainfall

data from the Live Oak station and neglecting runoff, which is assumed to be negligible

in this karst landscape. When rainfall sufficiently exceeded ET rates, water in the soil

Page 59: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

59

surpassed the moisture capacity, and recharge was recorded. For comparisons with

dissolution during spring reversals and bank infiltration, diffuse recharge was

considered only for the week preceding the spring reversal, 25 February to 3 March

2012. This week-long diffuse recharge value was multiplied by the area of the

Withlacoochee River drainage basin bound by the Pinetta Station on the north and the

Lee Station at the south. The drainage basin shapefiles used to determine the drainage

basin area were obtained from the Suwannee River Water Management District website

(http://www.srwmd.state.fl.us/index.aspx?NID=319). Maximum potential dissolution was

calculated by multiplying the basin-wide diffuse recharge (L) by the amount of calcite

that could be dissolved by intruding rain water (mmol/L), which was calculated using

rainfall chemistry, soil CO2 concentrations (see 3.4). These dissolution estimates were

later applied to the drainage basin using the yearly average rainfall to account for the

frequent nature of dissolution driven by rainfall in the study area.

Estimates of diffuse recharge use rainfall values and ET rates assumed to be

spread evenly across the entire drainage basin which is unlikely. The dissolutional

capability of rainwater reacting with soil CO2 will also vary with in the drainage basin

with soil depth, but these values are also assumed to be distributed uniformly across the

study area. We recognize these assumptions, along with those stated earlier, as

simplifications of a complex natural system, but data limitations make it impossible to

account for changes at every location. We, therefore, proceed with our calculations in

an attempt to better quantify dissolution in this system.

Page 60: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

60

3.3 Results

3.3.1 Rainfall and Recharge

On 3 March 2012, the Live Oak weather station recorded 37.3 mm of rainfall. In

the seven days preceding this event, the gauging station recorded 13.2 mm of rainfall,

for a total of about 50 mm. We estimate there was 20 mm of diffuse recharge during

this period. The drainage basin covers 7.8 x 108 m2 between the Pinetta and Lee

gauging stations and thus ~1.6 x 107 m3 of water recharged the aquifer, assuming

uniform rainfall across the basin and no runoff.

3.3.2 Hydrologic Responses

Pinetta and Madison Blue Spring river gauging stations show similar trends

through time, but the magnitude of changes in stage and discharge vary between the

gauging stations (Fig. 3-3). During the 2012 flood, from 2 to 9 March, the

Withlacoochee River rose 2.3 m at Pinetta, 2.4 m at Madison Blue, and 2.1 m at Lee.

Estimated discharge increased to a maximum of ~1.3 x 107 m3/day at Pinetta and ~1.2 x

107 m3/day at Lee (Fig. 3-4A). The gauging station at Madison Blue Spring run recorded

an increase in discharge up to ~3.4 x 105 m3/day until 2 March 2012, at which point

discharge declined rapidly, reaching a low of ~-2.6 x 105 m3/day during the reversal of

flow (Fig. 3-4A). These changes in river stage and discharge were sufficient to cause

intrusion to the spring that lasted 71 days before spring discharge returned to baseflow

water chemistry (Fig. 2-4).

3.3.3 Historic Record of High Water Events

The three methods used to estimate river intrusion were within an order of

magnitude of each other for this reversal. Gauging station data indicate ~7.2 x 105 m3,

whereas the groundwater flow model (Spellman 2012) and mixing model estimated ~2.2

Page 61: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

61

x 106 m3 and ~2 x 106 m3, respectively (Fig. 3-4B). We chose the hydrologic model

estimate for the remainder of the analyses (see 4-1).

Interactions between the river and surrounding aquifer differ in the two studied

river reaches (Pinetta-Madison Blue and Madison Blue-Lee). Immediately prior to the

2012 flood, the Pinetta-Madison Blue reach lost ~ 1.9 x 105 m3/day to the aquifer and

evapotranspiration. During the 2012 flood from 3 to 13 March 2012, the Pinetta-Madison

Blue reach lost an additional ~1.2 x 107 m3 over pre-event conditions, for a total input of

~1.4 x 107 m3 to the aquifer. Evapotranspiration for the area peaked at 3.3 mm/day

during the flood (FAWN), and since we are only concerned with evaporation, and not

transpiration, in the river, we consider any river losses, other than to the aquifer, to be

negligible. The Madison Blue-Lee reach is a gaining stream during baseflow and

received ~6.4 x 105 m3/day from the aquifer and surface runoff (Fig. 3-4C), but from 3 to

12 March 2012 the river gains slowed by ~4.6 x 106 m3 changing this river reach into a

losing river. Since these slowed gains are likely the result of river water entering the

aquifer, we add these decreases in aquifer discharge to the total river intrusion

calculation. From 3 to 22 March 2012, the river gains increased by ~1.3 x 107 m3 over

pre-flood levels. If we sum the losses from the Pinetta-Madison Blue river reach and

losses from Madison Blue-Lee river reach, excluding the Madison Blue reversal, we find

that ~1.4 x 107 m3 of additional river water above normal conditions entered the aquifer

during the 2012 flood.

Using the effective porosity and vertical and horizontal hydraulic conductivities of

the aquifer, we estimate the volume of matrix rock (e.g. Fig. 3-5) that the intruding river

water would occupy. We simplify the volume penetration shape assuming a cuboid

Page 62: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

62

which will be 1.5x greater in width than height because of the anisotropy of the

respective hydraulic conductivities (Williams and Kuniansky, 2015) (Fig. 3-1). River

bank intrusions in the Pinetta-Madison Blue reach would occupy 4 x 107 m3 of rock

matrix. The resulting surface expression of this value is 5.5 x 105 m2. The smaller

volume of intruded water to the Madison Blue –Lee reach would occupy 8 x 106 m3

generating a surface expression of 2.3 x 105 m2. Combining the two regions would

occupy a volume of 4.8 x 107 m3 of rock matrix with a surface expression of 7.8 x 105 m2

which will be used later in comparative denudation calculations. River water would

have intruded further into the matrix in some areas over others because conduits (i.e.

Madison Blue Cave System) would transport water farther away from the main channel

than matrix porosity. Nonetheless, these estimates provide an approximation of the

volume of the aquifer that could be occupied by intruded water.

The Madison Blue 2012 spring reversal is one of 124 reversal events that are

estimated to have occurred over the past 82 years, based on river discharge records

(Fig. 3-6), which equates to ~1.5 reversals/year. The 2012 reversal is in the first quartile

of river stage for all estimated spring reversals, which had no statistical outliers, which

are defined as values more than two standard deviations from the smallest or largest

recorded values.

3.3.4 Dissolution Estimates

The average composition of water intruding Madison Blue Spring indicates it was

capable of dissolving 0.17 mmol/L of calcite, assuming it reacted to equilibrium with

calcite. Multiplying this average dissolution potential by the hydrologic model estimate of

intruded river water volume (L) yields a total estimated dissolution during the spring

reversal of ~3.8 x 108 mmol of calcite, which based on a molar density of calcite of 2.71,

Page 63: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

63

with 30% porosity, represents a volume of ~1.4 x 107 cm3. Assuming the average

dissolution at Madison Blue Spring is similar to water loss to the river banks, an

additional ~2.4 x 109 mmol or ~9 x 107 cm3 of limestone would have dissolved along the

other portions of the river reaches.

Average rainfall chemistry in the region for March 2012 indicated rain water is

capable of dissolving ~0.42 mmol/L of calcite after soil CO2 concentrations of 10,000

ppm have reacted with infiltrating rain water. This concentration is frequently observed

(Karberg et al. 2005; Hasenmueller et al., 2015), and because of an increase in root

respiration following rainfall (Bouma and Bryla 2000), the soil pCO2 likely remains

elevated despite the fact that recharging water carries CO2 into the rock matrix. We

calculated that rainfall-driven diffuse recharge is capable of dissolving ~6.7 x 109 mmols

of calcite or ~2.6 x 108 cm3, but likely only dissolves 6 x 109 mmol or 2.3 x 108 cm3 (see

4.1). Combining dissolution values from all four sub-regions, we estimate a total of 8.8

x 109 mmol or 3.3 x 108 cm3 of calcite was dissolved during this event.

3.4 Discussion

The effects of spring reversals, river losses to the aquifer, and rainfall filtering

through the soil and rock matrix should impact the specific locations and extent of

dissolution across a landscape. Dissolution could be distributed widely across the

region, resulting in generally uniform denudation, or could be concentrated at point

sources, in conduits, or along river banks (Ritorto et al., 2009). The amount of

dissolution in each sub-region should also be modified by variations in the chemical

composition of the water. Spatial variations in dissolution should affect the regional

geomorphology and hydrogeology. Below, we report the relative amounts of dissolution

Page 64: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

64

that resulted from these differing mechanisms and discuss how they influenced

development of the modern landscape.

3.4.1 Dissolution Estimates

Chemical composition of river water differs from composition of rainfall,

suggesting a chemical control on the amount of possible dissolution. River water is

allogenic, largely originating as runoff from the Hawthorn Group in the northern portion

of the watershed (Fig. 3-2). This water contains high concentrations of organic matter

derived from wetlands perched on the confining unit. This organic matter produces CO2

when remineralized, but reaches pCO2 values less than those derived from infiltration

through the soil because of degassing in the river (Khadka et al., 2014). These

variations in composition lead to differences in the dissolution potential of the recharging

water. We discuss below the details of dissolution in each sub-region (Fig. 3-1).

During the spring reversal, hydrologic and mixing models indicate 1.4 x 107 cm3

and 1.25 x 107 cm3 of calcite was dissolved, based on the chemical composition of the

intruding water. This volume represents ~5% of the total dissolution caused by the

storm. Dissolution likely occurs on the conduit wall as well as in a “dissolution halo”

(Moore et al., 2010) in the matrix surrounding the conduits. Therefore, not all

dissolution enlarges the conduits, but most of the enlargement of the conduit appears to

result from intrusion of river water under current hydrologic conditions (Ezell et al., in

prep).

Loss of river water between the Pinetta and Madison Blue Spring gauging

stations was greater than pre-flood conditions, and the decrease in river gains between

Madison Blue Spring and the Lee gauging station reflect penetration of river water into

the aquifer before increased aquifer discharge. The water intruding the Madison Blue

Page 65: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

65

Cave system (~2.2 x 106 m3) represents ~44% of the reduced river water gains in the

Madison Blue-Lee reach (Fig. 3-4 B and C) and indicates that each river reach had

water penetrate into the aquifer despite transitioning from confined to unconfined

regions. Dissolution in this sub-region represents ~27% of the total dissolution caused

by the storm event. This dissolution is not distributed homogenously along both river

reaches, particularly considering the magnitude of water lost to Madison Blue Spring.

Other smaller conduits in the region likely also received some of the intruding water.

Nonetheless, assuming a homogenous penetration, based on vertical and horizontal

conductivities of river waters into the aquifer, allows us to estimate that the penetrated

portion of the matrix has a land surface expression area of ~7.8 x 105 m2 (Fig. 3-1).

Assuming the dissolution volume was evenly distributed over this area, a “denudation”

of 0.035 mm would have occurred by dissolution during the storm.

The chemical composition of recharged rainfall after reaction with soil CO2

indicates 0.42 mmol/L of calcite would dissolve, assuming a pCO2 of 10,000 ppm.

Diffuse recharge and estimated CO2 concentrations yield a total rainfall dissolution

value of 6.7 x 109 mmols of calcite or ~2.5 x 108 cm3, assuming the water reacted to

equilibrium. Because carbonate dissolution rates slow when waters approach

saturation with respect to calcite (Svensson and Dreybrodt 1992; White 2002), water is

unlikely to have completely equilibrated. Baseflow waters collected from the spring

were estimated to have recharged ~26 years prior to discharging according to CFC age

date estimates (Pers. Comm. Marie Kurz) and are still only 96-98% saturated with

respect to calcite. It is likely that at least a portion of diffuse recharge that occurred

during the storm discharged more rapidly than spring water flow during baseflow

Page 66: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

66

conditions, thereby preventing this diffuse recharge from reaching a baseflow discharge

saturation with respect to calcite. For comparison purposes, we assumed the water

reaches only 90% of saturation with respect to calcite. With this assumption, ~6 x 109

mmol or ~2.3 x 108 cm3 would have dissolved during the event, which is ~68% of the

storm-driven dissolution.

The magnitude of dissolution resulting from diffuse recharge is ~2.6 times greater

than river water dissolution (~9 x 107 cm3). Because this dissolution would be distributed

across the entire drainage basin, this value would convert to a denudation of ~0.0003

mm, which is <1% of the denudation caused by loss of river water to the aquifer.

Because this storm represented only ~3.7% of yearly rainfall and not all rainfall events

result in loss of water to the aquifer, the relative amounts of denudation caused by

recharge through the land surface and from loss of river water are probably more

similar. Assuming dissolution due to diffuse recharge is similar for all rainfall events, this

dissolution mechanism would represent denudation of 0.008 mm/yr, which is about 15%

of the denudation from loss of river water to the nearby aquifer during the year. This

denudation value is 10-50% of the 0.03 to 0.09 mm/yr previously estimated (Opdyke et

al. 1984; Adams et al. 2010), perhaps reflecting the additional dissolution from loss of

flood water.

The storm that we studied was not an especially strong storm event and did not

trigger an especially large spring reversal. The Madison Blue 2012 reversal ranked in

the first quartile of estimated reversals based on stage, and we speculate that bank

infiltration volumes associated with this event would correspond to stage and be in the

first quartile. While we do not know exactly how many spring reversals and bank

Page 67: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

67

infiltrations have occurred in the full ~80 year study record due to a lack of antecedent

moisture conditions/ground water level data, we can speculate that this storm triggered

a relatively small amount of dissolution compared to others in historical records (Fig. 3-

6).

3.4.2 Implications for Regional Geomorphology and Hydrology

Both loss of river water and diffuse recharge through the land surface (Fig. 3-1)

alter landscapes, but dissolution by these processes operates at different frequencies

and scales. Rainfall may occur without increasing river stage sufficiently to cause

spring intrusions or bank infiltration. Alternatively, rainfall outside the immediate

catchment area can flood the river, causing loss of water from the river to the aquifer

and dissolution, without having diffuse recharge to the aquifer through the land surface

(Brown, 2015). Although diffuse recharge associated with rainfall occurs more

frequently than spring reversals, loss of river water to the aquifer creates more focused

dissolution. Considering that intrusions occur ~1.5 times/yr (Fig. 3-6), this mechanism is

likely to be an important geomorphic control on the landscape.

Focused dissolution from loss of river water can enlarge previously created

voids. This dissolution pattern could produce the large spring/cave systems common to

north Florida. If regional denudation rates were higher, cave systems would be

exposed before they developed kilometers of passages (e.g. Madison Blue and

Peacock Springs), instead of opening only at rivers and sinkholes where dissolution is

more concentrated. Alternatively, if dissolution from loss of river water were greater

relative to dissolution from intrusion through the land surface, drainage would be

diverted underground, thereby limiting formation of rivers. This pattern of drainage is

seen on the Yucatan Peninsula, Mexico (Back and Hanshaw, 1970; Stringfield and

Page 68: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

68

LeGrand, 1976). Maintaining the proper ratio of rainfall dissolution to river water

dissolution has resulted in the hydrology and geomorphology in north Florida today,

which includes both large-conduit/spring systems and surface rivers.

3.5 Conclusions

The balance between focused dissolution from loss of river water and regional

denudation results in north Florida’s current geomorphology and hydrology, specifically

including long phreatic caves along with surface rivers. To determine these dissolution

patterns, we evaluated of how dissolution is distributed across a carbonate landscape

during a recharge event. We estimated total carbonate dissolution driven by this single

storm was 8.8 x 109 mmol or 3.3 x 108 cm3. River reversal into a single first magnitude

spring caused ~5% of the total dissolution. Loss of river water to the aquifer resulted in

27% of dissolution. The remainder of the dissolution (68%) was a consequence of

diffuse recharge across the land surface. Dissolution differences caused by these

mechanisms are a consequence of differences in recharge volume and degree of

undersaturation of the recharge water with respect to calcite. These differences in

recharge patterns and dissolution capability led to focused dissolution during the spring

reversal and river bank penetration and more dispersed dissolution during diffuse

recharge across the land surface. Focused dissolution from loss of river water is

responsible for enlarging conduits and spring systems which govern much of the

regional hydrology. In contrast, dissolution from diffuse recharge across the land

surface results in regional denudation. Comparative denudation values based on

dissolution from loss of river water at 0.05 mm/yr are similar to previously calculated

denudation range at 0.03-0.09 mm/yr, but this mode of dissolution is less frequent than

regional dissolution through rainfall, which created denudation values that are about one

Page 69: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

69

order of magnitude lower than the previously calculated denudation range at 0.008

mm/yr. This finding has implications for karst landscapes around the globe which

should vary according to local geology and climate; past and future landscape evolution

can now be better understood by estimating dissolution location and volumes based on

rainfall and drainage patterns.

Page 70: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

70

Figure 3-1. Karst terrain under normal and storm conditions. A) Karst terrain under

normal conditions showing a gaining river and spring additions to the river. B) Karst terrain under normal conditions showing a losing river and conduit loss from the river. C) Karst terrain under storm conditions showing higher river stage, higher water tables, and dissolution occurring due to rainfall, bank intrusion, and a spring reversal. Bank intrusion is 1.5x greater horizontally than vertically. Surface expression of rainfall infiltration is truncated on this figure. Modified from Gulley et al. (2013).

Page 71: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

71

Figure 3-2. Maps of the drainage basin and cave. A) Map of the Suwannee River Basin showing location of Madison Blue Springs cave systems and stage gauges on the Suwannee River. B) Map of Madison Blue Spring cave system in relation to the Withlacoochee River. Modified from Gulley et al. (2011).

Page 72: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

72

Figure 3-3. Stage data for Pinetta (dashed red line) and Madison Blue (black solid line) sites from 2002 through 2013. All stage data were corrected to the NGVD 29 datum and are shown as meters above sea level.

Page 73: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

73

Figure 3-4. River and spring responses to rainfall. A) Discharge at Pinetta, Madison Blue, and Lee during the 2012 flood. The increase in discharge is first and largest at Pinetta. The Madison Blue spring gauge shows negative discharge during the peak discharge at Pinetta and Lee. The falling limb of the hydrograph at Lee is elongated compared to Pinetta. B) Model-simulated (dotted line) and gauge-station-measured (solid line) discharge at the Madison Blue Spring gauging station. During the 2012 flood, the hydrologic model and USGS gauge data both show negative discharge, reflecting loss of water to the spring. The hydrologic model shows a greater negative discharge than the USGS data. C) River losses through time from Pinetta to Madison Blue river gauges and Madison Blue to Lee river gauges. The Pinetta to Madison Blue (green line) reach experiences river loss to the aquifer under normal flow conditions and losses increase during March 2012. The Madison Blue to Lee (black line) reach is normally a gaining river and briefly becomes a losing river before a rapid and short-lived increase in gains following the storm.

Page 74: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

74

Figure 3-5. Limestone river bank along the Withlacoochee River in north Florida. Photo

courtesy of Amy Brown.

Page 75: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

75

Figure 3-6. Pinetta Gauge Station stage data from 1932 to 2013 (black solid line) with calculated reversals (red dots).

Page 76: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

76

CHAPTER 4 BIOGEOCHEMICALLY DRIVEN CARBONATE DISSOLUTION AND FUTURE

CLIMATE IMPLICATIONS

4.1 Introduction

Because of the greenhouse effects of CO2, Earth’s temperature has fluctuated

between warm and cool periods that correspond to periods of elevated and depleted

atmospheric pCO2 (Shakun et al., 2012). Although strong ties exist between CO2 and

Earth’s temperature, the global carbon cycle remains poorly understood, especially with

regard to fluxes of CO2 to and from terrestrial carbonate reserves (Torres et al., 2014).

Because carbonate minerals represent the largest reservoir of C in the Earth,

dissolution and precipitation of these minerals should affect carbon cycling (Falkowski et

al. 2000; Martin et al., 2013; Liu and Dreybrodt, 2015).

Biogenically produced carbonate minerals form large carbonate platforms in

warm marine waters at low latitudes. These platforms dissolve when they react with

naturally produced acid, during which CO2 may be released to the atmosphere,

depending on the type of acid responsible for the dissolution (Torres et al., 2014). Two

types of acid, carbonic and sulfuric, are common in these settings, and both have been

identified as being capable of dissolution of carbonate platforms (Bottrell et al., 1991;

Gulley et al., 2013; Gulley et al., 2015; Stoessell et al., 1993). Net flux of CO2 to the

atmosphere from these dissolution reactions differs. When carbonic acid dissolves

limestone, two moles of bicarbonate are produced, with one mole of C from the

dissolution of solid carbonate and the other mole from atmospheric CO2 (Lerman et al.,

2007).

Page 77: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

77

𝐶𝑂2 + 𝐻2𝑂 ↔ 𝐻2𝐶𝑂3 (4-1)

𝐻2𝐶𝑂3 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 2𝐻𝐶𝑂3− (4-2)

The reverse of reactions 4-1 and 4-2 release CO2 back to the atmosphere that

previously had been sequestered as bicarbonate, and consequently, these coupled

reactions limit its impact on global carbon cycles over long periods of time (Berner et al.,

1983). In contrast, the coupling of oxidation of sulfide to sulfuric acid

𝐻2𝑆 + 2𝑂2 ↔ 𝐻2𝑆𝑂4 (4-3)

and the dissolution of carbonate minerals with that sulfuric acid

𝐻2𝑆𝑂4 + 𝐶𝑎𝐶𝑂3 ↔ 𝐶𝑎2+ + 𝐶𝑂2 + 𝐻2𝑂 + 𝑆𝑂42− (4-4)

generates a flux of CO2 to the environment that previously had been sequestered in the

solid phases.

Recent work suggests that burial of bicarbonate (produced in Reaction 4-2)

incorporated in organic matter represents a net sink of carbon from these systems (Liu

et al., 2011; Liu and Dreybrodt, 2015), at least on short time scales of a few hundred

years (De Vries et al. 2012). Understanding of the carbon cycle is further complicated

by detailed records of CO2 concentrations during the Cenozoic. In addition to carbonate

dissolution, silicate weathering can sequester atmospheric C, and during the Cenozoic,

uplift rates should have resulted in enough silicate weathering to remove all CO2 from

the atmosphere if a simultaneous source of CO2 did not exist. This proposed source

was CO2 released to the atmosphere as a result of carbonate dissolution by sulfuric

acid (Torres et al., 2014). With the understanding that carbonate mineral dissolution

may represent a sink, source, or have no impact on the global cycle, it becomes

Page 78: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

78

increasingly important to understand reaction mechanisms and how they relate to the

source of acid responsible for carbonate mineral dissolution.

The distributions and amounts of naturally occurring carbonic and sulfuric acid

vary widely. Carbonic acid is commonly formed from hydration of CO2 derived from

atmospheric CO2, plant respiration, and CO2 produced by microbial respiration during

the remineralization of organic matter (OM) (Adams and Swinnerton, 1937; Falkowski et

al., 1998). In contrast, sulfide distribution is more restricted because of its instability in

oxidizing systems, but it is common in and around marine systems because of the high

concentration of SO4 in seawater relative to other terminal electron acceptors (TEA) in

that environment (Stoessell et al., 1989, Schmitter-Soto, 2002). Sulfide is common at

haloclines that separate freshwater lenses that float isostatically on top of underlying

water with marine salinity (Socki, 1984; Stoessell et al., 1993). OM also accumulates at

the density interface of haloclines, where it is remineralized using the most energy-

efficient TEA. With O2 as a TEA, CO2 is released, but with SO4 as the less energetically

favorable, but more abundant TEA, both H2S and CO2 are formed. The result of these

reactions is that haloclines have low O2 concentrations and high concentrations of CO2

and H2S. Consequently, carbonate minerals may be dissolved at haloclines, which are

therefore ideal locations to study reaction mechanisms for carbonate mineral

dissolution.

Haloclines can be accessed through water-filled sinkholes; these features are

referred to as blueholes in the Bahamian carbonate platform and cenotes in the

Yucatan carbonate platform. Blueholes form by three processes: vadose dissolution

when sea levels are sufficiently low, bank margin failure, and progradational collapse of

Page 79: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

79

a subsurface void (Larson 2012). After blueholes form, they may expand at the land

surface to form a wide, circular surface expression (referred to here as the bowl) where

the conduit walls have been more extensively dissolved, and a narrower, more vertical

opening near the center of the structure (referred to here as the well). Processes that

could lead to this morphology have never been described, but we hypothesize that they

may relate to variations in water chemistry across the halocline.

Oxidation of H2S may occur through diffusion of atmospheric oxygen into the

water column or in situ production of O2 during photosynthesis. Diffusion of

atmospheric oxygen may restrict formation of sulfuric acid to near the water surface.

However, at greater depths in the water column, benthic plant photosynthesis would be

a more likely source of oxygen, provided that water depth and clarity permit sufficient

sunlight for primary production. Photosynthesis also consumes CO2, thereby increasing

pH. Diel patterns of changing pH values have been found in streams, rivers, and lakes

(Clark, 2002; Cicerone et al., 1999; de Montety et al., 2011; Kurz et al., 2013; Nimick et

al., 2003), but have never been measured at the interface of salt and fresh water of

carbonate platforms and evaluated in relationship to the potential for carbonate mineral

dissolution.

We estimated the relative magnitudes of dissolution caused by carbonic and

sulfuric acid to assess their potential for carbonate mineral dissolution and their roles in

CO2 transport to and from the atmosphere. This assessment was made using water

chemistry analyses logged at high frequency (1/min) and by direct sampling at lower

frequencies through tidal cycles and at various light intensities. The data were used to

address the role that tidal and solar radiation cycles have in controlling dissolution

Page 80: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

80

reactions. We used these data to evaluate how variations in dissolution of carbonate

minerals may influence the global carbon cycle.

4.2 Study Site and Methods

4.2.1 Study Site

This study focused on Inkwell Bluehole on San Salvador Island, The Bahamas

(24.1° N 74.48° W) (Fig. 4-1). San Salvador Island is the easternmost island in The

Bahamas and sits on an isolated carbonate platform. The island receives an average of

100.7 cm of annual rainfall (Shaklee, 1966). Approximately half of the island is covered

by lakes, most of which are either of marine salinity or hypersaline because of conduit

connections to the ocean and the negative water balance. Inkwell Bluehole, which is

located in the southwestern portion of the island (Fig. 4-1), is approximately circular in

shape and has a surface diameter of about 16 m and a depth of approximately 8 m (Fig.

4-2). No conduit has been found linking Inkwell directly to the ocean, but its tidal

amplitudes are usually within 15 cm of local ocean tide amplitude and its tides lag ocean

tides by only about 10 minutes, reflecting a high-permeability connection to the ocean

(Martin et al., 2012; Samson and Guilbeault, 2013). The water column is stratified, with

a halocline located at approximately 3 m depth. Overlying tannic surface water

attenuates light quickly as shown by Secchi disk readings of approximately 140 cm, but

water below 3 m is clear (Sampson and Gauilbeault, 2013).

4.2.2 Field Methods and Sample Analyses

Tidal fluctuations in the water level in Inkwell Bluehole were measured from 10 to

14 October 2012 by suspending a conductivity, temperature, and depth (CTD) sensor in

the bluehole water column at a fixed elevation through eight tidal cycles. Although the

absolute elevation of the logger is unknown, the difference from the mean elevation

Page 81: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

81

during the study provides tidal amplitudes and periodicity. The sampling period

occurred during three days leading up to a spring tide. During that time, three buoys

were secured in a transect across the top of the blue hole, with one buoy ~1 m from the

edge of the blue hole (nearshore), one in the center of the blue hole ~7 m from the

shore (center), and the third halfway between the other two, ~4 m from shore

(intermediate). Three logging instruments, including two InSitu Troll 9500s and one YSI

6600 multi-parameter logging instrument, were suspended from the buoys at 0.75 m

below the water surface. These instruments measured pH, temperature, specific

conductance (SpC), and dissolved oxygen (DO). The DO probe is optical, which

minimized interferences with sulfide and provided more accurate measurements at low

DO concentrations. Measurements were taken every minute throughout the experiment.

Since the buoys rose and fell with the tide, the instruments logged the water

composition at a fixed depth in the water column.

Water samples were collected three times from Inkwell Bluehole in October

2012, corresponding with the logging data, and twice in May 2012, when no logging

instruments were available. During both trips, samples were collected using a peristaltic

pump connected to flexible PVC tubing that extended from the shore to the center of the

bluehole in May 2012, and to each of the three buoys in October 2012. In May 2012,

samples were collected in a vertical profile through the water column while the PVC

tubing was hand-lowered at increments of 0.75 m at the center of the bluehole (Fig. 4-

3). During the October 2012 sampling, PVC tubing was suspended from each of the

three buoys at depths 2 m below the water surface. Samples were collected from just

prior to low tide to high tide on 11 October, from mid ebb tide to just after low tide on 12

Page 82: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

82

October, and from high tide to low tide on 14 October. Water was pumped into an open,

flow-through cell that held a sonde connected to a calibrated YSI 556MPS instrument

that measured specific conductance, DO, pH and temperature. Water was pumped

over the sonde for at least eight minutes until values stabilized. The DO concentrations

only show gross relationships with depth because the cup is open to the atmosphere,

and elevated sulfide concentrations can interfere with the sensor membrane, resulting in

an overestimate of DO concentrations.

Water samples for the measurement of major cations (Na+, K+, Mg2+, and Ca2+)

were filtered through a 0.45-µm trace-metal-grade high-capacity canister filter, collected

in acid-washed 20- mL plastic screw-top bottles, and preserved with trace-metal-grade

nitric acid to pH < 2. Water samples for measurements of anion concentrations (Cl-,

SO42-) were also filtered and collected unpreserved in new, but not acid-washed, 20- mL

plastic screw-top bottles. Water samples were collected in 60- mL pre-rinsed screw-top

plastic bottles and measured immediately in the field for hydrogen sulfide

concentrations using a HACH DR/890 Colorimeter using the Methylene Blue Method

(HACH Company 2013). Water samples for the measurement of dissolved inorganic

carbon (DIC) and δ 13C values of the DIC were collected in pre-rinsed 20- mL French

square vials and treated with HgCl2 to limit microbial alteration of the samples. Samples

were kept chilled on ice in the field and in refrigerators after returning to the lab at the

University of Florida, where major element, DIC concentrations, and δ 13C ratios were

measured.

Cation and anion values were measured on an automated Dionex DX500 Ion

Chromatograph. The samples had an average charge balance error of 1.6% with only

Page 83: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

83

one sample exceeding 5%. The sulfide concentrations were measured with an

accuracy of ±0.02 mg S/L. The DIC samples were measured by acidifying water

samples using an AutoMate Prep Device and the resultant CO2 was measured with a

UIC (Coulometrics) 5011 carbon coulometer. The method was standardized with

dissolved KHCO3 and yielded samples with an average accuracy of ±0.5%. Carbon

isotope ratios of DIC in water were measured with a Thermo Finnigan DeltaPlus XL

isotope ratio mass spectrometer with a GasBench inlet system. Water was injected into

septum-top vials that contained 0.5 mL phosphoric acid and were filled with helium.

This acidification released all DIC into the headspace and the gas mixture of CO2 and

He was sampled by the GasBench II and CO2 was separated by a GC column prior to

being measured on the mass spec. Carbon isotope measurements had a standard

deviation of 0.05.

4.2.3 Estimates of Potential Dissolution and Chemical Species

PHREEQc, a geochemical software package, was used to construct models to

estimate the production of CO2 and O2 needed to cause the observed decreases in pH

based on the reaction stoichiometry shown in reactions 4-1 and 4-3. The decrease in

pH was defined as the difference between the highest and lowest pH, i.e. between

shortly after the afternoon solar radiation maximum and when the sun rose (Fig. 4-4).

Water samples used to model these changes in pH were collected within a few hours of

the pH change (Fig. 4-4) and had pH values similar to those of the water column at the

start of the pH change as measured by the suspended loggers (Table 4-1). Grab

samples used in the pH models were taken from different times in the day to track the

changing water column chemistry through the day (Table 4-1).

Page 84: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

84

Water samples collected near the start of the pre-dawn pH decrease were

modeled by increasing the concentration of CO2 in PHREEQc until the pH of the

modeled water decreased by the same amount measured by the suspended loggers.

Similarly, water collected prior to the afternoon decrease in pH was modeled in

PHREEQc by reacting O2 and H2S until pH decreased to the observed range. Once pH

was decreased through either CO2 production or O2 and sulfide reactions, the water

was equilibrated with calcite in PHREEQc to assess its dissolution potential. The

differences in dissolution potential of the two acids represent their relative abilities to

dissolve calcite in the bluehole. These modeled changes in pH using sulfuric acid are

minima because consumption of CO2 and production of O2 during photosynthesis raises

pH, though we were unable to determine how much CO2 was taken up during

photosynthesis.

4.3 Results

Specific conductance shows haloclines at 3.75 m and 5.25 m below the water

surface in May 2012 at high and low tide, respectively, although the change in SpC

values at the halocline at high tide is sharper than at low tide (Fig. 4-3). Dissolved

organic carbon (DOC) concentrations were highest near the surface and decreased with

depth. DOC concentrations decreased more at high than low tide. Dissolved oxygen

percentages were highest at the surface during high tide and at 0.75 m depth during low

tide with nearly constant values of ~4% below 2.25 m depth, though some small portion

of this percentage may be an unquantifiable error resulting from the measurement

technique. Sulfide concentrations increased at the halocline at both high and low tides,

and the concentrations decreased sharply during high tide, but less sharply during low

tide (Fig. 4-3). Sulfide concentrations were near zero at the surface during both low and

Page 85: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

85

high tide and peaked at 3.75 meters below the water surface during low tide and 2.25

meters below the water surface during high tide.

Tidal amplitudes increased from 27 cm on 12 October to 65 cm on 14 October

and would have continued gaining amplitude until the spring tide, which occurred on 15

October (Fig. 4-4). Dissolved oxygen concentrations did not correspond to variations in

the water elevation and peaked only during late afternoon (Fig. 4-4). These values

cannot be compared with the DO depth profiles shown in Fig. 4-3 because of the

different measurement methods. For most of the record, oxygen saturation states were

below the detection limit, but display sharp maxima of up to 3.3% around 17:00 hrs

each day.

Water elevation varied inversely with several variables including pH, sulfide

concentration, δ 13C and DIC concentrations. pH decreased during high tide as shown

on both the in situ logger and the grab samples, with larger decreases occurring during

the day than at night. The in situ logger showed an average decrease in pH from high

to low tide of 0.24 during the day and 0.14 during the night (Fig. 4-4), whereas both the

in situ logger and the grab samples showed pH to be inversely correlated with tide

(Figs. 4-4 and 4-5). The grab samples show the same general trends in response to

tidal fluctuations at all three sites, but the variation in pH was diminished at the

nearshore site compared with other sites. Water elevation also correlated with sulfide

concentrations, which ranged from 1.3 mg/L during low tide to 17 mg/L during high tide

(Fig. 4-6). Water from the nearshore site contained more sulfide than the other sites at

low tide, occasionally by more than a factor of two, but the concentrations were similar

at all sites during high tide. Carbon isotope ratios showed the strongest inverse relation

Page 86: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

86

with water level in grab samples collected on 11 October, moving from low to high tide.

Values maintained the same general trend at all three sites (Fig. 4-7). All sites showed

similar trends of DIC concentrations through time relative to the tidal cycle, including a

plateau during high-tide conditions (Fig. 4-8).

Waters selected to model acid production and dissolution had pH values that

decreased by 0.07 to 0.18 pH units during the night and 0.2 to 0.31 pH units during the

day. The daytime decreases corresponded to increases in DO saturation states.

Assuming the nighttime pH changes resulted from increasing CO2 concentrations

caused by respiration, 0.2 to 0.4 mmol/L of CO2 was respired. Assuming the daytime

decrease in pH resulted only from oxidation of sulfide, between 0.5 and 0.8 mmol/L of

DO was reduced during the day (reaction 4-3). The DO would have been produced by

photosynthesis, which consumes CO2, and thus the observed decrease in pH must

represent only a minimum of sulfide oxidation. Prior to the production of CO2 at night,

the water had an SI of -0.02 and was capable of dissolving 0.017 mmol/L calcite. After

production of CO2, the SI decreased to -0.19, and the water was then capable of

dissolving 0.25 mmol/L. Prior to the oxidation of the sulfide, the water had an SI = -0.02

and was capable of dissolving 0.018 mmol/L calcite, but after the oxidation of sulfide the

SI decreased to -0.47 and the water was then capable of dissolving 0.62 mmol/L calcite.

4.4 Discussion

4.4.1 Relative Impact of Tidal and Solar Radiation Cycles on Dissolution Potential

Daily variations in both solar insolation and water level altered the saturation

state of the water with respect to calcite. Solar insolation is an important variable

through its control on photosynthesis and the consumption of CO2, production of

oxygen, and oxidation of sulfide to sulfuric acid. Tidal variations in water level influence

Page 87: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

87

the saturation state by raising and lowering the halocline relative to the water surface,

thereby bringing sulfide trapped at the halocline closer to atmospheric oxygen. The

amount the halocline moves through a tidal cycle relative to the water surface is a

function of the shape of the blue hole (Fig. 4-2). Movement of the halocline within the

water column, and diel cycles, change the pH of the water column at and above the

halocline. Daily changes in pH were also found in a United Kingdom stream where pH

varied from 7.68 to 8.48 during diel cycles (Spiro and Pentecost, 1991), and changes in

pH from 8.3 to 9.1, 6.9 to 7.3, and 8.1 to 8.3 were measured in three different Montana

streams by Nimick et al. (2003). All of these pH measurements showed lower values

during night than the day because of increasing CO2 concentrations, but this study

showed decreases in pH during the night and day. Whereas in both systems plants

respire CO2 leading to decreases in pH, this system also has sulfide, which oxidizes

and serves as an even more important control on pH.

In non-tidal freshwater systems, which contain no sulfide, pH values typically

reach minima at night as a result of respired CO2 and maxima during the day, as a

consequence of CO2 consumption during photosynthesis (Spiro and Pentecost, 1991;

Nimick et al., 2003). The variations in pH are particularly strong in freshwater systems

with abundant flora that are poorly buffered by carbonate minerals (Kurz et al., 2013),

and dissolution likely dampens decreases in pH in Inkwell associated with high tide (Fig.

4-4) through buffering before pH levels return to slightly basic conditions during low tide.

In contrast, the semi-diurnal variations of pH values at Inkwell Bluehole are impacted by

sulfur reactions as well as CO2 production and consumption, so that respiration of CO2

at night and oxidation of sulfide during the day regulate the water-column pH. Sulfuric

Page 88: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

88

acid is stronger than carbonic acid (pKH2SO4 = 2; pKH2CO3 = 10.3) (Krauskopf and

Bird, 2003), which indicates that at the same concentrations, sulfuric acid will generate

more H+ than carbonic acid and lower the pH more, leading to a greater amount of

carbonate dissolution.

During the three tidal cycles observed, the magnitude of the pH decrease during

the day correlates with the amount of O2 produced. Maximum recorded values indicate

a greater amount of sulfuric acid is produced through sulfide oxidation (0.8 mmol/L) than

carbonic acid by respiration (0.4 mmol/L). The daytime drop in pH from sulfuric acid

production would be even greater than that observed from CO2 consumption during

photosynthesis. Although we cannot know how much CO2 is taken up during

photosynthesis, theoretical calculations are possible. If even half of the greatest oxygen

production measured corresponded to an equal uptake of CO2, removing this uptake

would yield an additional 0.4 mmol/L of CO2 in the water. This is the same amount of

CO2 produced during the greatest nighttime pH drop recorded.

The maximum sulfide concentrations are 1.5 m closer to the water surface at

high tide than at low tide because of the upper portion of the water column spreading

across the wider bowl (Fig. 4-3D). As a result, sulfide at the halocline during high tide

has a greater potential for conversion to sulfuric acid by reaction with atmospheric and

photosynthetic oxygen. This observation is supported by Bottrell et al. (1991), who

noted that sulfide concentrations in cenotes are maximal at the halocline and are

oxidized to sulfate at shallower depths where carbonates are dissolved. Variations in δ

13C values and DIC concentrations also reflect fluctuations in the position of the

halocline. The δ 13C values are more negative during high than low tide, likely reflecting

Page 89: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

89

more δ 13C value control by remineralization of OM, generating CO2 byproducts than

with dissolution of CaCO3. The DIC concentrations vary with tide. This variation, the

inverse of δ 13C patterns, likely results from the same combination of solid carbonate

dissolving in water near the halocline and production of CO2 from remineralization. The

DIC concentrations and δ 13C values suggest that O2 produced through photosynthesis

and from the atmosphere is being used both as a TEA for microbes and oxidizer of H2S.

4.4.2 Implications for Platform Dissolution

The larger decrease in pH during the daytime than nighttime suggests that

sulfuric acid is the dominant driver of dissolution at Inkwell Bluehole. What remains

unknown is whether sulfuric acid could also represent an important cause of dissolution

in locations on carbonate platforms other than terrestrial blueholes. Dissolution occurs

throughout modern carbonate platforms as a result of multiple processes, including

mineral undersaturation caused by fresh and saltwater mixing (Mylroie and Carew,

1990; Plummer, 1975) and bacteria-produced CO2 from the remineralization of DOC in

the soil and rock matrix near the water table (Schwabe et al., 2008; Gulley et al., 2015).

To date, studies have not fully explored dissolution in carbonate platforms resulting from

generation of sulfuric acid outside of blueholes or cenotes.

Nonetheless, sulfuric acid should cause dissolution within the phreatic zone of

carbonate platforms when oxygen is available to drive the reaction. The question is

whether sulfide is generated in situ within the platform or if it is transported into the

matrix. Both possibilities exist. Water is exchanged between blueholes and the aquifer

matrix through tidal pumping (Martin et al., 2012). Tidal pumping results from

differences in phase and amplitude at the bluehole relative to the aquifer matrix as a

Page 90: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

90

result of differences in hydraulic conductivity of the aquifer and conduits connecting

blueholes with the ocean (Martin et al., 2012). This water exchange would allow sulfide

to enter the platform aquifer. Tidal variations of the water table will also allow

penetration of O2 from the water column as the vadose zone floods with rising tide. This

O2 would oxidize the sulfide in the aquifer, generating sulfuric acid (Brown et al., 2014).

Alternatively, DOC could be carried to the water table in the matrix, akin to the

input of particulate organic carbon in the blueholes. Similar to the porosity development

described in Schwabe et al. (2008) and Gulley et al. (2015), microbial remineralization

would generate byproducts leading to carbonate dissolution. In this case, microbial

remineralization would occur in anoxic environments at depth and the byproducts

generated would include H2S instead of solely CO2. When tidal pulses (near the shore)

or rainfall transport O2 into the aquifer, sulfuric acid would form and dissolve the

carbonate. The by-products would be transported away from the reaction site by the

regional groundwater flow, generating increased porosity. This hypothesis is supported

by sulfide measurements made in two well fields on San Salvador. One well in the

Linehole well field had a sulfide concentration of 1.7 mgS/L and one well in the Sandy

Point wellfield had a sulfide concentration of 8.5 mgS/L. These two wellfields are

located on opposite ends of the island, and whereas the maximum concentrations in the

wells vary, sulfide is likely present at depth across the entirety of the island, potentially

facilitating sulfuric acid dissolution of local rock.

4.4.3 Quantifying Dissolution Driven by Carbonic vs. Sulfuric Acid and C Flux to the Atmosphere

Carbonate mineral dissolution by sulfuric and carbonic acid results in different

potential fluxes of C to the atmosphere (reactions 4-1 – 4-4), indicating that

Page 91: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

91

differentiating the relative amount of dissolution by these different acids is important to

determine C exchange between the lithosphere and atmosphere. We differentiate

these fluxes by evaluating the potential amount of calcite dissolution from each acid.

Approximately 0.25 mmol/L calcite would dissolve if water with a pH acidified through

the addition of CO2 at dawn was equilibrated with calcite. In contrast, approximately

0.62 mmol/L calcite would dissolve, due to oxidation of sulfide decreasing pH in the

afternoon, or approximately 2.5 times more calcite than the estimated carbonic acid

dissolution. Dissolution by sulfuric acid is more than twice the dissolution by carbonic

acid though maximum acid production of sulfuric acid was only two times more than

carbonic acid (0.8 vrs. 0.4 mmol/L, respectively). This discrepancy in dissolution

volumes potentially stems from the tendency toward back-reactions associated with

bicarbonate in solutions and changing pH, whereas sulfate is more stable. Each of

these reaction products (Reactions 4-2 and 4-4) offsets the positive charge associated

with calcium ions, and if the offsetting charges of bicarbonate are removed, the solution

may not be able to dissolve as much calcite. The difference in sulfuric and carbonic

acid dissolution potential also extends to atmospheric carbon interactions; sulfuric acid

could add up to 0.8 mmol C/L/tidal cycle to the atmosphere if the waters reached

complete equilibrium with calcite, whereas carbonic acid dissolution would not source C

to the atmosphere.

The morphology of the bluehole also suggests that sulfuric acid drives most of its

dissolution. When Inkwell was originally formed, it likely did not have the nearly circular

surface expression seen today, but instead would have had an irregularly shaped

surface opening caused by dissolution along a preferential flow path in the vadose zone

Page 92: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

92

or collapse of a subsurface void. The flat bowl that occurs at the level of the high-tide

halocline (Fig. 4-2) suggests that the majority of Inkwell expansion through dissolution

has occurred in this portion of the water column, long enough to give the bluehole its

distinctive shape today. We do not know how long Inkwell has been dissolving, but the

morphology suggests that the majority of its formation occurred when sea levels were

high enough that it was at least partially water-filled.

Dissolution patterns in Inkwell under current sea levels have been established,

but sea levels vary, and here we speculate on the impact of sea levels rising and falling

from their current position. When sea levels rise, a mix of fresh and salt water will

persist in Inkwell until the saltwater completely inundates the feature after an ~6m rise

from its current position. The number of land-based blueholes will continue to decline

as the ocean overtops the platform containing them. Aerial exposure of the carbonate

platform will continue shrinking until the entire island will theoretically be overtopped.

Rising sea levels will limit sulfuric acid dissolution because of the greater distance

between the photic zone, which produces O2, and the majority of the carbonate

platform, until the process stops completely when no land is left above the sea. A

shrinking platform exposure will mean less dissolution driven by carbonic acid and

terrestrial OM production, but shallow sea productivity will likely increase as the majority

of the carbonate platform is covered by a shallow ocean.

The surface area of San Salvador Island, The Bahamas, is smaller than the

isolated carbonate platform it sits atop, but the Bahamas Banks contain several other

Bahamian islands and many would be connected by land if sea levels dropped 25 m

from present levels. During falling sea levels more blueholes would be present on the

Page 93: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

93

larger exposed platform, which should increase surface water/groundwater interactions.

These increased interactions would allow O2 to penetrate into the platform matrix and

react with hydrogen sulfide to form sulfuric acid (Reaction 4-3), increasing this type of

dissolution across the platform. The lower sea levels would also shift the freshwater/salt

water interface lower with respect to the carbonate platform and away from the current

surface exposure of San Salvador. This shift would remove the sulfur source from the

current exposure of San Salvador which would severely limit any further dissolution of

the current island by sulfuric acid. Dissolution driven by carbonic acid should increase

with more sub-aerial reaction sites on the larger exposed platform, which would also

support more terrestrial OM production. With steep-sided edges of the platform

exposed, shallow sea productivity would be limited during lower sea levels.

In addition to the impacts of sea level variation on island systems, sources and

sinks of atmospheric carbon will also vary. During high sea levels there is diminished

sourcing of C to the atmosphere from sulfuric acid dissolution, but there is also less

carbonic acid dissolution and terrestrial OM burial, which interact to serve as a C sink.

During low sea levels, C sources and sinks will increase with surface water/groundwater

interaction. Although it is difficult to predict the impact of changing sea levels on C

cycling in island settings, we can make general assumptions about the expected trends,

and in this paper, we identified more sulfuric acid dissolution than carbonic acid

dissolution at surface water/groundwater interaction sites. If at lower sea levels more of

these interaction sites are present, lower sea levels should generate a greater source of

C to the atmosphere via the processes described in this paper. This greater source of

C to the atmosphere may serve as a negative feedback loop through warming the

Page 94: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

94

atmosphere as a greenhouse gas, which would ultimately raise sea levels (Shakun et

al., 2012). More extensive research is needed in locations other than surficial openings

to the water table, but given the data presented in this paper, carbonate dissolution

should not be discounted when attempting to balance the global C cycle.

4.5 Conclusions

Tidal and diel cycles impact the potential for carbonate dissolution in Inkwell

Bluehole, The Bahamas, by influencing carbonic and sulfuric acid concentrations.

Dissolution potential resulting from the production of sulfuric acid is more than twice that

of carbonic acid. The geomorphology of Inkwell supports sulfuric acid as the dominant

dissolution driver with a wide upper dissolved zone that corresponds to the high sulfide

halocline. Carbonate dissolution driven by carbonic acid is understood to occur across

carbonate platforms, but we propose that sulfuric acid may also dissolve carbonate

platforms at depth. Dissolution driven by sulfuric acid represents a net C flux to the

atmosphere, and carbonic acid dissolution tied to plant burial represents a C sink.

Sulfuric acid dissolution of carbonate is greater than that of carbonic acid, which means

that Inkwell serves as a net source of C to the atmosphere. This study identifies

changes in dissolution potential that are greatest during the day; in most studies

dissolution potential is greatest at night, and this dissolution pattern in Inkwell reflects

the presence of sulfuric acid, which is not present in most continental study sites.

Dissolution in island systems on carbonate platforms needs to be better quantified in

locations other than surficial openings to the groundwater. These dissolution

assessments would better define the role of dissolution on a carbonate platform as a net

C sources or sinks to the atmosphere, but this role would be subject to change with

variations in sea level. Studies in blueholes, such as Inkwell, provide windows to

Page 95: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

95

aquifers in carbonate platforms that allow a better understanding of processes that

dissolve carbonate platforms in tropical settings around the world.

Page 96: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

96

Table 4-1. Water chemistry and parameters for grab samples used to model changes in water column pH recorded by loggers.

Inkwell Nearshore

YSI Readings Ions (mmol/L)

Date and Time Temp C SpC

uS/cm DO %

pH Cl- SO42- Ca2+ Na+ Mg2+ K+

Sulfide mgS/L

Alkalinity HCO3

10/11/12 14:27 28.36 17363 3.8 6.93 157.5251 7.2377 5.7748 141.2618 15.2957 3.0952 10.75 6.20

10/12/12 8:47 28.2 17068 4.40 6.99 153.5319 7.0493 5.5895 134.8571 14.7919 2.9504 10 6.20

10/12/12 13:45 29.02 15953 3.80 7.02 142.7423 6.6667 5.2278 123.9348 13.741 2.7184 6.3 5.43

10/14/12 8:10 27.77 19229 2.5 6.88 176.8657 8.0489 5.7787 158.4971 17.3446 3.4238 17 6.19

10/14/12 13:10 28.92 15957 3 7.09 136.5011 6.4208 5.0555 117.3618 13.1235 2.5547 5.4 5.06

Page 97: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

97

Figure 4-1. Map of The Bahamas and San Salvador. A) Map showing location of The Bahamas in relation to Florida with San Salvador Island boxed. B) San Salvador Island, indicating the location of Inkwell Bluehole, Linehole wellfield, and Sandy Point wellfield. Modified from Gulley et al. (2015).

Page 98: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

98

Figure 4-2. Conceptual model of a cross section of Inkwell Bluehole during high and low tide. The halocline is located further down the well during low tide and at the top of the well during high tide. The fresh water is thicker during low tide and thinner during high tide. Red circles represent hydrogen sulfide distribution in the water column. Figure modified from Samson and Guilbeault (2013).

Page 99: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

99

Figure 4-3. Depth profiles collected on 6 May 2012 showing A) specific conductance, B) dissolved organic carbon concentrations, C) dissolved oxygen, and D) sulfide at high and low tide. Collection depths are relative to the high tide water elevation.

Page 100: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

100

Figure 4-4. Dissolved oxygen percent saturation, water depth, and pH plotted through time at the nearshore site in Inkwell. Gray bars indicate night, white bars indicate daytime, and yellow bars indicate periods when grab samples were collected.

Page 101: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

101

Figure 4-5. pH values and variation from mean water level near shore, intermediate,

and center sites in Inkwell occurring on A) 11 Oct 2012 B) 12 Oct 2012 and C) 14 Oct 2012.

Page 102: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

102

Figure 4-6 Sulfide concentrations and variation from mean water level for near shore, intermediate, and center sites in Inkwell occurring on A) 11 Oct 2012 B) 12 Oct 2012 and C) 14 Oct 2012.

Page 103: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

103

Figure 4-7. δ 13C values and variation from mean water level for near shore, intermediate, and center sites in Inkwell occurring on A) 11 Oct 2012 B) 12 Oct 2012 and C) 14 Oct 2012.

Page 104: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

104

Figure 4-8 Dissolved inorganic carbon concentrations and variation from mean water

level for near shore, intermediate, and center sites in Inkwell occurring on A) 11 Oct 2012 B) 12 Oct 2012 and C) 14 Oct 2012.

Page 105: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

105

CHAPTER 5 SUMMARY

Carbonate dissolution has been studied in both laboratory and field settings.

Dissolution rates have been calculated under a variety of conditions, including flowing

and stagnant waters, changing pCO2, a range of pH, in the presence of different ions,

and several more. This research, however, was one of the first investigations to link

natural time series data sets and observed biogeochemistry of waters to study

carbonate dissolution rates and volumes.

Carbonate dissolution in karst settings is driven by rainfall that reacts with the

land surface and can force river water into river banks and spring systems. The

dissolution driven by rainfall was found to be greater than dissolution driven by river

water intruding into the river banks and springs. River water dissolution also occurred

less frequently than rainfall dissolution, but river water dissolution is more intensely

concentrated than rainfall-driven dissolution, which is spread more evenly over the

drainage basin. The balance between the two dissolution drivers is important because

it results in north Florida’s current geomorphology and hydrology. If rainfall dissolution,

considered denudation here, were much greater, conduits would not reach the size

seen today and subsurface hydrology would be affected. If river water dissolution were

much greater, virtually all landscape drainage in north Florida might go underground

resulting in a loss of surface rivers. The balance seen today is the reason north Florida

currently has long phreatic caves, while retaining surface rivers.

Dissolution driven by river water forced into spring systems can be up to 17

orders of magnitude faster than rates seen at baseflow. This makes dissolution in these

systems episodic. Approximately 80% of dissolution during the spring reversals

Page 106: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

106

occurred in the first 20% of the flood, before rates slowed almost to baseflow rates near

the end of the intrusion. Initial dissolution was driven by undersaturation with respect to

calcite of the intruding river water. Undersaturation is prolonged during the reversal by

remineralization of dissolved organic carbon that intrudes with the flood water. Diffuse

recharge of rainfall through the land surface is slightly undersaturated by the time it

reaches the conduits, but results in little dissolution of the conduits because of

exponential slowing of dissolution rates near saturation. At the beginning of a flood,

dissolution alone adds Ca2+ to intruding river water, but when the flood recession

begins, simple mixing adds more Ca2+ to conduit waters and is primarily responsible for

slowing dissolution rates.

Carbonate dissolution can also be driven by tidal and diel cycles in blueholes.

The biogeochemical shifts introduced by these cycles alter H2CO3 and H2SO4

concentrations, which in turn control carbonate dissolution rates and volume.

Dissolution potential resulting from the production of H2SO4 was more than twice the

dissolution driven by H2CO3. This pattern is the reverse of what is seen in most karst

systems with biological control. Normally, the majority of dissolution occurs during the

night when CO2 is respired by plants, but in this system, dissolution occurs primarily

during the day when plants are photosynthesizing. This dissolution pattern is further

supported by bluehole morphology, which has a wide upper bowl formation near the

surface as a result of increased dissolution. The lower edge of this bowl aligns well with

the lowest position of the halocline, which is the source of the H2S reacting to form

H2SO4. Dissolution driven by H2SO4 represents a net C flux to the atmosphere, unlike

dissolution by carbonic acid. Blueholes provide windows to aquifers in carbonate

Page 107: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

107

platforms that allow a better understanding of processes that dissolve carbonate

platforms in tropical settings around the world.

Page 108: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

108

LIST OF REFERENCES

Adams, C.S., Swinnerton, A.C., 1937. Solubility of limestone. Eos, Transactions American Geophysical Union 18, 504–508.

Adams, P.N., Opdyke, N.D., Jaeger, J.M., 2010. Isostatic uplift driven by karstification and sea-level oscillation: Modeling landscape evolution in north Florida. Geology 38, 531–534.

Albéric, P., Lepiller, M., 1998. Oxydation de la matière organique dans un système hydrologique karstique alimenté par des pertes fluviales (Loiret, France) oxidation of organic matter in a karstic hydrologic unit supplied through stream sinks (Loiret, France). Water Research 32, 2051–2064.

Albury, N., 2010. Sawmill Sink, Abaco, Bahamas: A multi-environment fossil preservation system, in: 2010 GSA Denver Annual Meeting.

Back, W., Hanshaw, B.B., 1970. Comparison of chemical hydrogeology of the carbonate peninsulas of Florida and Yucatan. Journal of Hydrology 10, 330–368.

Berner, R.A., Caldeira, K., 1997. The need for mass balance and feedback in the geochemical carbon cycle. Geology 25, 955–956.

Bottrell, S.H., Smart, P.L., Whitaker, F., Raiswell, R., 1991. Geochemistry and isotope systematics of sulphur in the mixing zone of Bahamian blue holes. Applied Geochemistry 6, 97–103.

Bouma, T.J., Bryla, D.R., 2000. On the assessment of root and soil respiration for soils of different textures: interactions with soil moisture contents and soil CO2 concentrations. Plant and Soil 227, 215–221.

Brown, A.L., Martin, J.B., Screaton, E.J., Ezell, J.E., Spellman, P., Gulley, J., 2014. Bank storage in karst aquifers: the impact of temporary intrusion of river water on carbonate dissolution and trace metal mobility. Chemical Geology 385, 56–69.

Budd, D.A., Vacher, H.L., 2004. Matrix permeability of the confined Floridan Aquifer, Florida, USA. Hydrogeology Journal 12, 531–549.

Buhmann, D., Dreybrodt, W., 1985a. The kinetics of calcite dissolution and precipitation in geologically relevant situations of karst areas: 1. Open system. Chemical geology 48, 189–211.

Buhmann, D., Dreybrodt, W., 1985b. The kinetics of calcite dissolution and precipitation in geologically relevant situations of karst areas: 2. Closed system. Chemical Geology 53, 109–124.

Page 109: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

109

Cantrell, K.J., Serkiz, S.M., Perdue, E.M., 1990. Evaluation of acid neutralizing capacity data for solutions containing natural organic acids. Geochimica et Cosmochimica Acta 54, 1247–1254.

Carew, J.L., Mylroie, J.E., 1995. Depositional model and stratigraphy for the Quaternary geology of the Bahama Islands. Geological Society of America.

Chen, X., Chen, X., 2003. Stream water infiltration, bank storage, and storage zone changes due to stream-stage fluctuations. Journal of Hydrology 280, 246–264.

Cicerone, D.S., Stewart, A.J. & Roh, Y., 1999. Diel cycles in calcite production and dissolution in a eutrophic basin. Environmental toxicology and chemistry, 18(10), pp.2169–2177.

Clarke, S.J., 2002. Vegetation growth in rivers: influences upon sediment and nutrient dynamics. Progress in Physical Geography, 26(2), p.159.

Covington, M.D., Gulley, J.D., Gabrovšek, F., 2015. Natural variations in calcite dissolution rates in streams: Controls, implications, and open questions. Geophysical Research Letters 42, 2836–2843.

Culver, D.C., White, W.B., 2005. Encyclopedia of caves. Elsevier Amsterdam (The Netherlands).

De Montety, V., Martin, J.B., Cohen, M.J., Foster, C., Kurz, M.J., 2011. Influence of diel biogeochemical cycles on carbonate equilibrium in a karst river. Chemical Geology 283, 31–43.

DeVries, T., Primeau, F., Deutsch, C., 2012. The sequestration efficiency of the biological pump. Geophysical Research Letters 39.

Dove, P.M., Platt, F.M., 1996. Compatible real-time rates of mineral dissolution by Atomic Force Microscopy (AFM). Chemical Geology 127, 331–338.

Dreybrodt, W., 1990. The role of dissolution kinetics in the development of karst aquifers in limestone: A model simulation of karst evolution. The Journal of Geology 639–655.

Dreybrodt, W., 1998. Principles of karst evolution from initiation to maturity and their relation to physics and chemistry. Science Press, Beijing.

Dreybrodt, W., Gabrovsek, F., Gabrovsek, F., 2002. Basic processes and mechanisms. Evolution of Karst: From Prekarst to Cessation 115–154.

Ezell, J., Martin, J. B., Brown, A.L., Gulley, J., 2013. Tidal effects on the chemical composition of water within a bluehole: a possible control on their morphologic evolution, in: 2013 GSA Annual Meeting in Denver.

Page 110: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

110

Falkowski, P.G., Barber, R.T., Smetacek, V., 1998. Biogeochemical controls and feedbacks on ocean primary production. Science 281, 200–206.

Falkowski, P., R. J. Scholes, E. Boyle, J. Canadell, D. Canfield, J. Elser, N. Gruber, K. Hibbard, P. Halgberg & S. Linder, 2000: The global carbon cycle: a test of our knowledge of earth as a system. - science 290, 291-296.

Farrell, M.D. and Upchurch, S.B. 2004. Development of Madison Blue Spring-Based MFL Technical Report. Suwannee River Water Management District Department of Water Resources Report

Florea, L.J., Vacher, H.L., 2006. Springflow hydrographs: eogenetic vs. telogenetic karst. Ground Water 44, 352–361.

Florea, L.J., Vacher, H.L., Donahue, B., Naar, D., 2007. Quaternary cave levels in peninsular Florida. Quaternary Science Reviews 26, 1344–1361.

Ford, D.C., Palmer, A.N., White, W.B., 1988. Landform development; karst. Geological Society of America.

Ford, D., Williams, P.D., 2013. Karst hydrogeology and geomorphology. John Wiley & Sons.

Franz, R., Franz, S.E., 2009. A New Fossil Land Tortoise in the Genus Chelonoidis (testudines: Testudinidae) from the Northern Bahamas: With an Osteological Assessment of Other Neotropical Tortoises. University of Florida.

Hach Company, DR/890 Colorimeter Procedures Manual. Methylene Blue Method, Method 8131. 2013. 48470-22, printed in USA

Gèze, B., 1987. Les mésaventures des sources de l’Estavelle et de l’Inversac en Languedoc Méditerranéen. International Journal of Speleoleology 16, 101–109.

Gischler, E., Shinn, E.A., Oschmann, W., Fiebig, J., Buster, N.A., 2008. A 1500-year Holocene Caribbean climate archive from the Blue Hole, Lighthouse reef, Belize. Journal of Coastal Research 1495–1505.

Granger, D.E., Fabel, D., Palmer, A.N., 2001. Pliocene- Pleistocene incision of the Green River, Kentucky, determined from radioactive decay of cosmogenic 26Al and 10Be in Mammoth Cave sediments. Geological Society of America Bulletin 113, 825–836.

Gulley, J.D., Martin, J.B., Moore, P.J., Brown, A., Spellman, P.D., Ezell, J., 2015. Heterogeneous distributions of CO2 may be more important for dissolution and karstification in coastal eogenetic limestone than mixing dissolution. Earth Surface Processes and Landforms.

Page 111: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

111

Gulley, J.D., Martin, J.B., Moore, P.J., Murphy, J., 2013. Formation of phreatic caves in an eogenetic karst aquifer by CO2 enrichment at lower water tables and subsequent flooding by sea level rise. Earth Surface Processes and Landforms 38, 1210–1224.

Gulley, J., Martin, J.B., Screaton, E.J., Moore, P.J., 2011. River reversals into karst springs: A model for cave enlargement in eogenetic karst aquifers. Geological Society of America Bulletin 123, 457–467.

Hach Company, DR/890 Colorimeter Procedures Manual. Methylene Blue Method, Method 8131. 2013. 48470-22, printed in USA

Hasenmueller, E.A., Jin, L., Stinchcomb, G.E., Lin, H., Brantley, S.L., Kaye, J.P., 2015. Topographic controls on the depth distribution of soil CO2 in a small temperate watershed. Applied Geochemistry 63, 58–69.

Hemond, H.F., 1990. Acid neutralizing capacity, alkalinity, and acid-base status of natural waters containing organic acids. Environmental Science & Technology 24, 1486–1489.

Hercod, D.J., Brady, P.V., Gregory, R.T., 1998. Catchment-scale coupling between pyrite oxidation and calcite weathering. Chemical geology 151, 259–276.

Hoorn, C., Wesselingh, F.P., Ter Steege, H., Bermudez, M.A., Mora, A., Sevink, J., Sanmartín, I., Sanchez-Meseguer, A., Anderson, C.L., Figueiredo, J.P., others, 2010. Amazonia through time: Andean uplift, climate change, landscape evolution, and biodiversity. science 330, 927–931.

Huggett, R., 2007. Fundamentals of geomorphology. Routledge.

Jacobson, R.L., Langmuir, D., 1974. Controls on the quality variations of some carbonate spring waters. Journal of Hydrology 23, 247–265.

Jennings, J.N., Jennings, J.N., 1985. Karst geomorphology. Blackwell Oxford.

Johnson, D.B., Hallberg, K.B., 2005. Acid mine drainage remediation options: a review. Science of the total environment 338, 3–14.

Jørgensen, B.B., 1980. Seasonal oxygen depletion in the bottom waters of a Danish fjord and its effect on the benthic community. Oikos 68–76.

Jouzel, J., Barkov, N.I., Barnola, J.M., Bender, M., Chappellaz, J., Genthon, C., Kotlyakov, V.M., Lipenkov, V., Lorius, C., Petit, J., and others, 1993. Extending the Vostok ice-core record of palaeoclimate to the penultimate glacial period. Nature 364, 407–412.

Page 112: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

112

Karberg, N.J., Pregitzer, K.S., King, J.S., Friend, A.L., Wood, J.R., 2005. Soil carbon dioxide partial pressure and dissolved inorganic carbonate chemistry under elevated carbon dioxide and ozone. Oecologia 142, 296–306.

Kaufmann, G., Dreybrodt, W., 2007. Calcite dissolution kinetics in the system CaCO3–H2O–CO2 at high undersaturation. Geochimica et cosmochimica acta 71, 1398–1410.

Khadka, M.B., Martin, J.B., Jin, J., 2014. Transport of dissolved carbon and CO2 degassing from a river system in a mixed silicate and carbonate catchment. Journal of Hydrology 513, 391–402.

Krauskopf, K.B., Bird, D.K., 1967. Introduction to geochemistry. McGraw-Hill New York.

Kurz, M.J., de Montety, V., Martin, J.B., Cohen, M.J., Foster, C.R., 2013. Controls on diel metal cycles in a biologically productive carbonate-dominated river. Chemical Geology 358, 61–74.

Larson, E.B., 2012. Modeling blue hole development in The Bahamas, in: 2012 GSA Annual Meeting in Charlotte.

Lerman, A., Wu, L., Mackenzie, F.T., 2007. CO2 and H2SO4 consumption in weathering and material transport to the ocean, and their role in the global carbon balance. Marine Chemistry 106, 326–350.

Liu, Z., Dreybrodt, W., 2015. Significance of the carbon sink produced by H2O–carbonate–CO2–aquatic phototroph interaction on land. Science Bulletin 60, 182–191.

Liu, Z., Dreybrodt, W., Liu, H., 2011. Atmospheric CO2 sink: Silicate weathering or carbonate weathering? Applied Geochemistry 26, S292–S294.

Mann, G.K., 2002. Karst landscape evolution. Evolution of Karst: From Prekarst to Cessation 243.

Martin, J.B., Dean, R.W., 2001. Exchange of water between conduits and matrix in the Floridan aquifer. Chemical Geology 179, 145–165.

Martin, J.B., Gulley, J., Spellman, P., 2012. Tidal pumping of water between Bahamian blue holes, aquifers, and the ocean. Journal of Hydrology 416, 28–38.

Martin, J.B., Brown, A., Ezell, J., 2013. Do carbonate karst terrains affect the global carbon cycle?/Ali kraska obmocja na karbonatih vplivajo na global no krozenje oglijika? Acta Carsologica 42, 187.

Meinzer, O.E., 1927. Large springs in the United States. US Govt. Print. Off.

Page 113: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

113

Moore, P.J., Martin, J.B., Screaton, E.J., Neuhoff, P.S., 2010. Conduit enlargement in an eogenetic karst aquifer. Journal of Hydrology 393, 143–155.

Mylroie, J.E., Carew, J.L., 1990. The flank margin model for dissolution cave development in carbonate platforms. Earth Surface Processes and Landforms 15, 413–424.

Mylroie, J.E., Carew, J.L., 1995. Karst development on carbonate islands. Ch.3 M 63: Unconformities and porosity in carbonate strata. etd. Budd, D.A., Saller, A.H., Harris, P.M., AAPG.

Mylroie, J.E., Carew, J.L., Moore, A.I., 1995. Blue holes: Definition and genesis. Carbonates and Evaporites 10, 225–233.

NCDC. 2015.Rainfall data. http://www.ncdc.noaa.gov/oa/climate/normals/usnormals.html. National Climate Data Center. Accessed July 7, 2015.

Nimick, D.A., Gammons, C.H., Cleasby, T.E., Madison, J.P., Skaar, D., Brick, C.M., 2003. Diel cycles in dissolved metal concentrations in streams: occurrence and possible causes. Water Resources Research 39.

Opdyke, N.D., Spangler, D.P., Smith, D.L., Jones, D.S., Lindquist, R.C., 1984. Origin of the epeirogenic uplift of Pliocene-Pleistocene beach ridges in Florida and development of the Florida karst. Geology 12, 226–228.

Palmer, A.N., 1991. Origin and morphology of limestone caves. Geological Society of America Bulletin 103, 1–21.

Parkhurst, D.L., Appelo, C.A.J., others, 1999. User’s guide to PHREEQC (Version 2): A computer program for speciation, batch-reaction, one-dimensional transport, and inverse geochemical calculations.

Pitty, A.F., 1968. The scale and significance of solutional loss from the limestone tract of the southern Pennines. Proceedings of the Geologists’ Association 79, 153–IN6.

Plummer, L.N., 1975. Mixing of sea water with calcium carbonate ground water. Geological Society of America Memoirs 142, 219–236.

Plummer, L.N., Wigley, T.M.L., Parkhurst, D.L., 1978. The kinetics of calcite dissolution in CO2-water systems at 5 degrees to 60 degrees C and 0.0 to 1.0 atm CO2. American Journal of Science 278, 179–216.

Ramanathan, V., 1981. The role of ocean-atmosphere interactions in the CO2 climate problem. Journal of the Atmospheric Sciences 38, 918–930.

Page 114: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

114

Ritorto, M., Screaton, E.J., Martin, J.B., Moore, P.J., 2009. Relative importance and chemical effects of diffuse and focused recharge in an eogenetic karst aquifer: an example from the unconfined upper Floridan aquifer, USA. Hydrogeology Journal 17, 1687–1698.

Sampson, J., Guilbeault, K., 2013. Baseline physicochemical investigations on waters from three blue holes, San Salvador Island, Bahamas. Studia UBB, Geologia 58, 11–19.

Schmitter-Soto, J.J., Comín, F.A., Escobar-Briones, E., Herrera-Silveira, J., Alcocer, J., Suárez-Morales, E., Elías-Gutiérrez, M., Díaz-Arce, V., Marín, L.E., Steinich, B., 2002. Hydrogeochemical and biological characteristics of cenotes in the Yucatan Peninsula (SE Mexico). Hydrobiologia 467, 215–228.

Schwabe, S.J., Herbert, R.A., Carew, J.L., 2008. A hypothesis for biogenic cave formation: a study conducted in the Bahamas, in: Proceedings of the Thirteenth Symposium on the Geology of the Bahamas and Other Carbonate Regions. Gerace Research Centre, San Salvador. pp. 141–152.

Shaklee, R.V., 1966. Weather and Climate, San Salvador Island, Bahamas. Bahamian Field Station, Limited.

Shakun, J.D., Clark, P.U., He, F., Marcott, S.A., Mix, A.C., Liu, Z., Otto-Bliesner, B., Schmittner, A., Bard, E., 2012. Global warming preceded by increasing carbon dioxide concentrations during the last deglaciation. Nature 484, 49–54.

Shiraki, R., Rock, P.A., Casey, W.H., 2000. Dissolution kinetics of calcite in 0.1 M NaCl solution at room temperature: an atomic force microscopic (AFM) study. Aquatic Geochemistry 6, 87–108.

Shuster, E.T., White, W.B., 1971. Seasonal fluctuations in the chemistry of lime-stone springs: A possible means for characterizing carbonate aquifers. Journal of Hydrology 14, 93–128.

Socki, R.A., 1984. A chemical and isotopic study of groundwater from Northwestern Yucatan, Mexico. M. S. thesis. Department of Geology, Northern Illinois University, DeKalb, Illinois.

Spellman, P., 2012. River Losses at a Karst Escarpment During Normal Flow and Flood Conditions and Implications for Carbonate Weathering. University of Florida.

Spence, J., Telmer, K., 2005. The role of sulfur in chemical weathering and atmospheric

CO2 fluxes: evidence from major ions, δ 13C DIC, and δ34S SO4 in rivers of the Canadian Cordillera. Geochimica et Cosmochimica Acta 69, 5441–5458.

Page 115: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

115

Spiro, B., Pentecost, A., 1991. One day in the life of a stream—a diurnal inorganic carbon mass balance for a travertine-depositing stream (waterfall beck, Yorkshire). Geomicrobiology Journal 9, 1–11.

SRWMD. http://www.srwmd.state.fl.us/. Suwannee River Water Management District.

Steadman, D.W., Franz, R., Morgan, G.S., Albury, N.A., Kakuk, B., Broad, K., Franz, S.E., Tinker, K., Pateman, M.P., Lott, T.A., 2007. Exceptionally well preserved late Quaternary plant and vertebrate fossils from a blue hole on Abaco, The Bahamas. Proceedings of the National Academy of Sciences 104, 19897–19902.

Steadman, D.W., Morris, J.R., Wright, N.A., 2013. A new species of Late Pleistocene rail (Aves: Rallidae) from Abaco, the Bahamas. Paleontological Journal 47, 1355–1364.

Stoessell, R.K., Moore, Y.H., Coke, J.G., 1993. The occurrence and effect of sulfate reduction and sulfide oxidation on coastal limestone dissolution in Yucatan cenotes. Groundwater 31, 566–575.

Stoessell, R.K., Ward, W.C., Ford, B.H., Schuffert, J.D., 1989. Water chemistry and CaCO3 dissolution in the saline part of an open-flow mixing zone, coastal Yucatan Peninsula, Mexico. Geological Society of America Bulletin 101, 159–169.

Stringfield, V.T., LeGrand, H.E., 1976. Karst hydrology of Northern Yucatan Peninsula, Mexico.

Svensson, U., Dreybrodt, W., 1992. Dissolution kinetics of natural calcite minerals in CO2-water systems approaching calcite equilibrium. Chemical Geology 100, 129–145.

Szymczak, P., Ladd, A.J., 2011. The initial stages of cave formation: Beyond the one-dimensional paradigm. Earth and Planetary Science Letters 301, 424–432.

Torres, M.A., West, A.J., Li, G., 2014. Sulphide oxidation and carbonate dissolution as a source of CO2 over geological timescales. Nature 507, 346–349.

Vacher, H.L., Mylroie, J.E., 2002. Eogenetic karst from the perspective of an equivalent porous medium. Carbonates and Evaporites 17, 182–196.

White, W.B., 1988. Geomorphology and hydrology of karst terrains. Oxford university press New York.

White, W.B., 2002. Karst hydrology: recent developments and open questions. Engineering geology 65, 85–105.

White, W.B. 2005. Hydrogeology of karst aquifers. In Encyclopedia of Caves, ed. D.C. Culver and W.B. White, 293–300. Amsterdam: Elsevier Academic Press.

Page 116: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

116

White, W.B., 2009. The evolution of Appalachian fluviokarst: competition between stream erosion, cave development, surface denudation, and tectonic uplift. Journal of Cave and Karst Studies 71, 159–167.

Williams, L.J., Kuniansky, E.L., 2015. Revised hydrogeologic framework of the Floridan aquifer system in Florida and parts of Georgia, Alabama, and South Carolina. US Geological Survey.

Page 117: To My Familyufdcimages.uflib.ufl.edu/UF/E0/04/95/19/00001/EZELL_J.pdf · flooding of karst springs ..... 20 2.1 introduction ... 3 relative amounts of dissolution in carbonate terrains

117

BIOGRAPHICAL SKETCH

John Eric Ezell was born in Tifton, Georgia. He moved to Mississippi and spent

most of his time in the woods or on a sports field. Along the way he met a few people

who encouraged him before finishing high school including Ms. Williams who had a little

extra time to help a forgetful elementary student, Ms. Huber who showed that some

teachers really were friends, and Dr. Padgett who embodied everything a teacher could

hope to be. He next attended Mississippi State University for a BS in Forestry, a BS in

Environmental Geosciences, and a MS in Geology. While at Mississippi State he

worked with one of the smartest minimum wage crews in history, had the chance to

learn from and joke with some of his favorite professors (Drs. Ezell, Grebner, McNeal,

and Mylroie come to mind), participate in a 36 hour sampling extravaganza, and

receive, possibly his greatest accomplishment, a shirt for winning the intramural softball

league. He then moved to Florida where he received his PhD in Geology and met a lot

of great friends. He, very slowly, got smarter about his research sites progressing from

briar covered clear cuts, to muddy estuaries, and finally to crystal clear springs. He had

many great instructors and friends along the way. He actively looks forward to spending

more time with family and finding a nice comfortable cave.


Recommended