<<

Interpretation of Whether Incision Rates in Appalachian Karst Reflect Long-term

Downcutting toward a Surface Versus Subsurface

A thesis presented to

the faculty of

the College of Arts and Sciences of Ohio University

In partial fulfillment

of the requirements for the degree

Master of Science

Holly A. Fitzgibbon

November 2010

© 2010 Holly A Fitzgibbon. All Rights Reserved.

2 This thesis entitled

Interpretation of Whether Incision Rates in Appalachian Karst Reflect Long-term

Downcutting toward a Surface Versus Subsurface Base Level

by

HOLLY A. FITZGIBBON

has been approved for

the Department of Geological Sciences

and the College of Arts and Sciences by

Gregory S. Springer

Associate Professor of Geological Sciences

Benjamin M. Ogles

Dean, College of Arts and Sciences 3 Abstract

FITZGIBBON, HOLLY A., M.S., November 2010, Geological Sciences

Interpretation of Whether Incision Rates in Appalachian Karst Reflect Long-term

Downcutting toward a Surface Versus Subsurface Base Level (154 pp).

Director of Thesis: Gregory S. Springer

Evidence of landscape evolution is recorded in packages and

cave passage morphologies found in Buckeye Creek Cave (BCC), West Virginia.

Sediments and cave passages have been age-dated using U/Th and paleomagnetism so

that long-term incision rates could be determined. The oldest cave passages are

>773000 old and the youngest passage is still forming. Incision rates calculated for

higher, older cave passages are between ≤27 and ≤71 meters/million years, which is

consistent with rates elsewhere in Appalachia. However, whether BCC incision rates

reflect long-term incision or karst processes is ambiguous because passages may

have been adjusted to a subsurface base level as opposed to a surface and valley.

This would explain a low incision rate (≤27m/Ma) obtained from atop thick in

the second lowest passage, 16.3m above the modern stream, and why comparatively little

incision has occurred since the original passage floor was excavated.

Keywords: , Climate, Incision, Karst, Paleomagnetism, Cave, Speleothem

Approved:

Gregory S. Springer

Associate Professor of Geological Sciences 4 Acknowledgements

I wish to thank the entire Department of Geological Sciences here at Ohio

University, but especially I wish to thank my advisor and committee members, Dr.

Springer, Dr. Lee, and Dr. Nadon. Each provided me with help and insight when I needed it. Especially Dr. Springer, who spent a cold, wet night with me in Buckeye Creek Cave when it flooded. I would also like to thank Cheri Sheets and Tim Grubb for everything. I would also like to thank the other graduate students in our department, my friends who helped me stay motivated. Of course I am eternally grateful for the Geological Sciences

ALUMNI research grant I received.

I would like to thank Deanna Potkanowicz, Ph.D., for your wonderful insights. I would also like to thank Thomas Carney, my high school Earth Sciences professor. I owe him everything because without him I never would have chosen this field.

Finally, I would like to thank my family and my fiancé Patrick. Mom, Dad, you gave me a wonderful childhood and early adulthood. You paid for all of my schooling. I hope I’ve made you proud. Patrick, there aren’t enough words to express how thankful I am, you were here, you kept me sane, worked a job you disliked, and were always there to take me in early or pick me up late. Thank you so much.

Thank you all. 5 Page

Table of Contents

Abstract ...... 3

Acknowledgements ...... 4

List of Tables ...... 8

List of Figures ...... 9

1.0 Introduction ...... 10

1.1 Previous studies in Buckeye Creek Cave ...... 10

1.2 Local Climate Changes ...... 12

1.3 Landform Responses to Climate Changes ...... 16

1.4 Cave Sediment and Speleothems ...... 17

1.5 Denudation ...... 19

1.5.1 Chemical Weathering ...... 19

1.5.2 Physical Weathering and Incision ...... 22

1.5.3 ...... 24

1.5.3.1 Stream Power ...... 24

1.5.3.2 Shear Stress ...... 26

1.6 Purpose ...... 28

1.7 Research Questions and Hypothesis ...... 29

2.0 Study Area ...... 33

2.1 ...... 33

2.2 Landforms ...... 39

2.2.1 Drainage Boundaries ...... 39 6 2.2.2 Base Level Position ...... 39

2.2.2.1 Creek ...... 41

2.2.3 Passage Morphologies ...... 45

2.2.4 Sediment Packages and Speleothems ...... 46

3.0 Methodology ...... 47

3.1 Surveying ...... 47

3.2 Sieving ...... 48

3.3 Paleomagnetism ...... 49

3.4 Uranium-thorium ...... 54

3.5 Carbon ...... 59

4.0 Results ...... 64

4.1 Elevation Data ...... 65

4.2 Age Data ...... 65

4.3 Incision Rates ...... 66

4.4 Stope’s Top Sediments ...... 66

4.5 Snail Shell Site ...... 67

4.6 Sieving Results ...... 68

4.7 Wolman Count ...... 69

5.0 Discussion ...... 74

5.1 Comparison of incision rates ...... 74

5.2 Spring Creek ...... 81

6.0 Conclusions ...... 85

7.0 References ...... 86 7 8.0 Appendix ...... 101

8.1 Appendix A: Survey Data ...... 101

8.2 Appendix B: Sieving Results ...... 117

8.3 Appendix C: Graphs of grain size (mm) vs. % finer ...... 141

8.4 Appendix D: Sieving Statistics ...... 153

8.5 Appendix E: Wolman Count Statistics ...... 154 8 Page

List of Tables

Table 1. List of (MIS) for the past 420,000 years...... 15

Table 2. Representative isotopic data from various organic matter ...... 62

Table 3. Table summarizing incision rates calculated for this study...... 64

Table 4. Summary of the sieving results showing the 16th, 50th, and 84th percentiles...... 72

Table 5. Wolman count performed on the sandstone cobbles at the Snail ...... 73

Table 6. Incision rates as determined from age and elevation data from ...... 79

9 Page

List of Figures

Figure 1. Stalagmite BCC-002 δ13C and δ18O values for the last 6,000 years ...... 14

Figure 2. Relationship of speleothems to external climate-drivers ...... 21

Figure 3. Location of Greenbrier County and the Greenbrier watershed ...... 35

Figure 4. Location of study site within the Greenbrier watershed ...... 36

Figure 5. Generalized stratigraphic section of the Greenbrier Group ...... 37

Figure 6. Location of study area. BC is Buckeye Creek, UE is the upstream ...... 38

Figure 7. Plan view of Buckeye Creek Cave. The black arrows indicate ...... 43

Figure 8. Cross-section of Buckeye Creek Cave along Q-q showing the four ...... 43

Figure 9. Map of Buckeye Creek Cave showing the upstream and downstream ...... 44

Figure 10. Massive cobble fill atop of laminated silts, , and gravels ...... 70

Figure 11. Hjulstrom curves velocities required for , transportation ...... 71

Figure 12. Cartoon showing subsurface vs. surface channels of Spring ...... 84 10 1.0 Introduction

1.1 Previous studies in Buckeye Creek Cave

Since the 1939 description of Buckeye Creek Cave (BCC) by Paul H. Price and

E.T. Heck, extensive exploration and surveying has taken place. Although BCC has been well studied with regard to its sedimentological and speleothem records, it has also been studied with regard to methods of channel incision (Springer and Wohl, 2002; Springer et al., 2003), hydraulics (Springer, 2004), Native American land disturbances

(Springer et al., 2010), and cave biology (Schneider and Culver, 2004).

Springer and Wohl (2002) observed the walls and floors of Buckeye Creek Cave in order to study sculpted forms, which occur during channel erosion. Caused by and dissolution, these forms often occur along planes of weakness as the result of recirculating vortices. The sizes of the sculpted forms can be used in models, which accommodate the effects of basin area, channel gradient, and rock resistance to determine vortex size and structure as well as efficiency. These and other processes gradually lower or incise the cave floor toward local base level.

Buckeye Creek flows through sandstones and shales prior to entering the limestones of BCC. In a study of correlations between reach-scale channel hydraulics and substrate resistance Springer et al. (2003) determined that unit stream power and shear stress are highest atop the shale and decline in the downstream direction across the limestones through which Buckeye Creek Cave is formed. Springer et al. (2003) supports the hypothesis that can adjust channel geometries and hydraulics to substrate characteristics while maintaining concave profiles. Thus Buckeye Creek can grade to local base level (presently Spring Creek). 11 Springer (2004) modeled energy losses in segments generated by large-scale flow separation associated with expansions and bends. At the Watergate, a constriction in BCC with a low ceiling and relatively calm waters, measurements of gravel size, water depth, and cross sections were taken to be used in a model which would provide expected energy losses at such a constriction. The model yielded results in agreement with qualitative expectations whereby energy losses were low except in the constriction where energy losses were great. The model was for closed-conduit or pipe-full flow and demonstrated that channel erosion need not be limited to open channel flow.

Using age stalagmites from BCC, Springer et al. (2008) demonstrated that six centennial-scale droughts occurred during the last five North ice- events during solar minima. The droughts provide evidence for solar forcing of east-central North American droughts and for solar modulation of mid-continent climates. In addition, Springer et al. (2010) showed that where the record does not correlate with the five North Atlantic Ocean ice-rafting events it can be linked to periods of time when Native Americans inhabited the region. Through burning and clearing, the

Native Americans could have caused vegetation changes above and around the cave.

These vegetation changes are also preserved in the stalagmites in addition to the ice- rafting events. A speleothem from BCC (BCC-002) was used to develop a stable isotope record of climate changes (Figure 1). In addition, -deposited cave sediments were used to construct a history of aggradation, incision and morphological change in the surface channel. Between collection of the stalagmites and the clastic cave sediments values for δ13C and δ18O were obtained. The stable isotopic ratios are different

environmental proxies; δ13C reveals vegetation changes while δ18O reveals temperature 12 and precipitation changes within region. In addition to vegetation changes, the cave sediments themselves can be used to infer past hydrological relationships between the cave and base level (Springer et al., 2009).

1.2 Local Climate Changes

Climate changes throughout the have been extensively studied. These changes are preserved in vegetation records, speleothem records, limnologic records, and fluvial sediments (Laird et al., 1996; Denniston et al., 1999; Baker et al., 2001). Climate changes are recorded in the literature as a numbered Marine Isotope Stages (MIS) which increases further back in time. The even numbered stages are glacial periods and the odd numbered stages are periods (Table 1).

Climate is dominated by cycles with periods near 23,000, 41,000 and 100,000 years (Imbrie et al., 1992). The 23,000 and 41,000 cycles have influenced virtually every part of the global climate system for more than 500,000 years. These cycles are linear, continuous responses to orbitally-driven changes in the radiation budget of the

Earth. The precession, obliquity, and eccentricity of the Earth are also climatically important through their control of the latitudinal and seasonal distribution of incident solar radiation. These factors either set the phase of oscillations that are driven internally or drive the major climate cycles externally (Imbrie et al., 1992). Importantly, change in relative wetness affect the ability of a stream to incise, so the Pleistocene climate changes are one possible driver of incision rates in Appalachian karst (Granger et al., 2001).

Glacial/interglacial cycles can be broken down by various time scales. At the scale of 100,000s of years is a glacial/interglacial sequence, 10,000s of years is one cold- warm cycle, 1,000s of years is one instability phase, and at the scale of 100s of years are 13 lower order climatic changes (Vandenberghe, 1995). Therefore, in the last 300,000 years there have been three major glacial/interglacial cycles (Lisiecki and Raymo, 2007). Three cool and arid climate episodes (300,000 – 250,000; 190,000 – 130,000; 100,000 – 4,000 years ago) are recorded in the most up-to-date BCC stalagmite record (Springer, G.S., personal communication). When compared with ice volume reconstructions, these cool and arid periods do correlate with high ice volumes on continents.

Climatic fluctuations have been extensively studied with the result that climate reconstructions of reveal cool, but warming temperatures in the early

Holocene (10,000 to 7,000 BP). The early to mid Holocene (7,000 to 4,000 years B.P.) shows significant warming known as the Hypsithermal (Laird et al., 1996; Denniston et al., 1999). After the Hypsithermal (7,000 to 4,000 years B.P.), temperatures dropped significantly (Denniston et al., 1999). The Late Holocene is characterized as a time of cool, but fluctuating, humid temperate climates (Mackay et al., 2005). Although did not extend into southeastern West Virginia, glacial maximums resulted in a much cooler and drier climate (Leigh, 2008) causing periglaciation both locally and regionally

(Springer, 2002; Nelson et al., 2007). Periglaciation refers to the mobilization and creation of sediment by frost and ice action, often in association with permafrost. These changes also caused the Greenbrier River, near BCC, to aggrade (fill with sediment) and later incise those sediments (Springer et al., 2009). This once again demonstrates the importance of as a control on incision.

14

Figure 1. Stalagmite BCC-002 δ13C and δ18O values for the last 6,000 years. From White 2007.

15 Table 1. List of Marine Isotope Stages (MIS) for the past 420,000 years.

MIS Name Climate Beginning End Reference(s)

Holocene 1 Interglacial Interglacial 12,000 - Kines et al., 2001

2 Wisconsin Glacial 25,000 12,000 Kines et al., 2001

3 Wisconsin Interglacial 59, 000 25,000 Lea et al., 2000 Lea et al., 2000; Kines 4 Wisconsin Glacial 74, 000 59,000 et al., 2001 Potter and Lambeck, 5a Wisconsin Interglacial 84,000 72,000 2003

5b Wisconsin Interglacial 94,000 85,000 Kines et al., 2001

5c Wisconsin Interglacial 108, 000 94,000 Kines et al., 2001

5d Wisconsin Interglacial 115,000 104,000 Shackelton et al., 2003 Potter and Lambeck, 5e Interglacial 128,000 118,000 2003 Fairbanks and 6 Glacial 180,000 125,000 Matthews, 1978

7 Pre-Illinoian Interglacial 245,000 190,000 Robinson et al., 2002

8 Pre-Illinoian Glacial 334,000 245,000 Stirling et al., 2001 Stirling et al., 2001; 9 Pre-Illinoian Interglacial 340,000 334,000 Lea et al., 2000 Droxler and Farrell, 10 Pre-Illinoian Glacial 360,000 360,000 2000; Lea et al., 2000 Droxler and Farrell, 11 Pre-Illinoian Interglacial 420,000 360,000 2000

16 1.3 Landform Responses to Climate Changes

Climatic fluctuations, such as transitions from glacial to interglacial periods drive temperature and precipitation changes which directly affects stream hydrology as a result

(Ely, 1997). Thus, channel networks are exceptionally sensitive to changes in climate

(Tucker and Slingerland, 1997). Geomorphic evolution in karst terrains is preserved in three main forms; cave passage morphologies, sediment packages (discussed in the next section), and passage development and abandonment resulting from climate-driven cave stream piracies (Johnson and Gomez, 1994; Granger et al., 2001).

On the surface as well as in caves, climatically induced, rapid stream incision results in the formation of high, but narrow stream (Palmer, 1987; Granger et al.,

2001). Rapid incision can also cause phreatic caves to be abandoned in favor of narrow vadose canyons (Granger et al., 2001). Phreatic cave passages are permanently flooded to their ceilings and pressure head drives flow in a manner analogous to pipes. Vadose cave passages are partially air-filled and governed by the same driving forces as those found in surface streams. On the surface, river aggradation produces alluvial terraces and underground aggradation partially or completely infills caves with sediment (Granger et al., 2001; Springer et al., 2009). For example, if an increase in storm frequency or magnitude raises overland flow shear stresses, then and sediment influx will increase leading to aggradation (Tucker and Slingerland, 1997).

Stream deposits, often interpreted as evidence of glacial/interglacial cycles, are often preserved in caves (Granger et al., 2001). As such, cave sediments can be records of past climatic warming and cooling events and the effects these changes had on landscapes (Granger et al., 2001; Springer et al., 2009). Dating of cave sediments can 17 constrain the timing of cave stream infilling (aggradation) for comparison to glacial climates recorded in stalagmites (e.g., Granger et al., 2001; Springer et al., 2009). Thus, sediments can be used determine rates of channel infilling and downcutting (incision), which can then be used to estimate erosion rates on surrounding landscapes at the times of and during particular climate regimes (Palmer, 1987; Tebbens et al., 1999;

Reusser et al., 2004; Leigh 2008).

Sedimentation rates and sediment nature is strongly influenced by climate which in turn is linked to effects on vegetation, weathering, and transport mechanisms. During humid or wet climate periods when there are increased precipitation rates, surface sediments are more easily eroded and hydrologic systems are more capable of transporting sands and larger grain sizes (Wendland, 1996; Wilby et al., 1997). Thus, under ideal conditions, changes in climate are can be inferred based on an analysis of the grain sizes and sedimentary characteristics such as crossbedding or imbrication of fluvial deposits (White, 2007)

By understanding how climate change impacts the fluvial landscape it is possible to predict the impact of future changes in climate and land use (Tucker and Slingerland,

1997; Thompson, 1999; Dyurgerov et al., 2000; Leigh, 2008). Additionally, understanding this impact can be used for deciphering the stratigraphic and geomorphic record of climate change (Tucker and Slingerland, 1997).

1.4 Cave Sediment and Speleothems

The deposition of cave sediments as well as the formation of speleothems is an extensively studied topic. In eastern North America, large temperature decreases during glacial periods are associated with major decreases in precipitation (Springer et al., 18 2008), which lowers sediment transport capacity, but increases sediment supply resulting in aggradation (Granger et al., 2001; Springer, 2005; Springer et al., 2009).

Sediment will accumulate in stream passages if driving forces are incapable of moving all the sediment supplied to a channel and this can be recorded as sediment banks on channel margins, thick deposits beneath streambeds, and terrace-like deposits in abandoned cave passages (Ford and Williams, 2007). These deposits may result from climate changes where it becomes more arid or when excess sediment is supplied to caves or nearby by streams, glaciers, or landsliding (Vandenberghe, 1995;

Springer et al., 2009). If climate becomes so cool that soils freeze and become permafrost, periglaciation may occur and the vegetation that was holding the soil in place will be replaced by sparser plant communities or bare ground. This can release abundant soil into Appalachian streams and rivers and commonly results in aggradation (Eaton et al., 2003). Buckeye Creek Cave contains multiple levels (tiers) and several significant sediment packages. Dating of these sediments may establish a link between sediment aggradation, climate changes, and hillslope erosion.

The most up-to-date Buckeye Creek Cave stalagmite δ13C and δ18O time series

spans the last 300,000 and is correlated with global ice volumes (Springer, G.S., personal

communication). Interestingly, where the cause of river aggradation is uncertain,

comparisons with previous studies often attribute them to general climatic cooling and

the growth of glaciers (Granger et al., 2001). The most recent large-scale cooling episode

prevailed from 100ka to 20ka and is well recorded by regional periglacial landforms

(Bender et al., 1994; Petit et al., 1999; Nelson et al., 2007) and BCC stalagmites. 19 Speleothems are exceptionally useful for determining ages of stream passages due to the fact that speleothem growth is episodic and can begin to accumulate in cave passages anytime after they are drained of groundwater. Climate change often initiates and stops speleothem growth (Stock et al., 2005b). Speleothems can record climate variability, rainfall variability, atmospheric circulation changes, and vegetation response

(Figure 2) and are ideal materials for precise U-series dating (McDermott, 2004).

1.5 Denudation

Denudation is the rate at which a terrain is lowered by erosion. Denudation rates have been determined for many basins around the world. However, of interest to this study are corrosion-associated denudation rates of basins near Buckeye Creek Cave:

Dickson Spring Basin, Walters’ Spring Basin, and Cold Spring Basin. Ogden (1982) determined karst denudation rates for these specific basins by using spring water hardness to determine the amount of dissolved carbonate. Geologic time scales of decalcification can be calculated with knowledge of Ca2+ concentrations in water (Drever, 1997). Ogden

(1982) found that for Dickson Spring Basin the dissolution rate was 22.6±5

mm/1000years. For Walters’ Spring Basin it was 22.3±5 mm/1000years and for Cold

Spring Basin it was 19.0±5 mm/1000years. Although the basin areas varied widely (from

2.4 to 64 km2) the denudation rates were relatively similar (Ogden, 1982).

1.5.1 Chemical Weathering

Chemical weathering includes those processes by which rocks are decomposed by

chemical alteration of their minerals, resulting in a change in the composition of

weathered materials (Price and Heck, 1939; Bland and Rolls, 1998; Monroe and

Wicander, 2006). The relationship between chemical weathering and climate is complex 20 (Kump et al., 2000) because weathering rates depend on factors such as the supply of water, its residence time in the soil and initial pH, the mineralogy of the rocks exposed, the abundance of organic acids, the temperature of soil solution, and the reactive surface area of the minerals, (Drever, 1997; Kump et al., 2000). Plants typically increase the CO2 content of a soil through indirect and direct processes (Drever, 1997; Kump et al., 2000) and they incorporate both chemical and physical effects into weathering (Kump et al.,

2000). The chemical effects are related to the production of organic acids, which affect soil water pH, and can enhance dissolution (Kump et al., 2000). Through the process of photosynthesis, carbon dioxide in the atmosphere is accumulated into organic matter, which is then decomposed in the soil (Drever, 1997; Yoshimura et al., 2001) and either emitted as CO2 into the atmosphere and or dissolved in soils by rainwater in the course of

its flow from the surface to underground (Yoshimura et al., 2001).

In a short time scale, local climate and CO2, a proxy for carbonic acid, movement

via groundwater can closely be related to karst landform development, including incision

because the solubility of carbonate rocks is high chemically aggressive waters. Calcite,

such as largely composes limestones and marbles, is practically insoluble in pure water,

however, it rapidly dissolves if a small amount of acid is present (Bland and Rolls, 1998;

Monroe and Wicander, 2006). One way to make water acidic is by dissociating the ions

+ of carbonic acid as follows: H2O + CO2 ↔ H2CO3 ↔ H + HCO3¯. Once an acidic

solution is present, calcite rapidly dissolves according to the reaction: CaCO3 + H2O +

2+ CO2 ↔ Ca + 2HCO3¯ (Drever, 1997; Yoshimura et al., 2001; Monroe and Wicander,

2006). The rapid dissolution of calcite leads easily to depletion of calcite from sediments,

soils, and rocks, including those in stream beds. 21

Temperature changes in the atmosphere atmosphere and hydrological cycle Atmosphere (composition and circulation e.g. NAO, Changes in SOI) Solar Inputs

Sulphur Aerosols Windblown Dust Cloud

Cooling Rainout Rayleigh Fractionation Volcanic Activity Runoff Evaporation 18O Enrichment Cryosphere: Sea ice, ice sheets, Alpine cave system Glaciers Sulphur 18O Depletion Aerosols Ice Sheet

Human Influences Land uses Biosphere

Sea Ice Soil - Biosphere Interaction Ocean (changes in circulation (changes in response to e.g. NADW, sea level, etc) precipitation, temperature etc)

Allogenic Water

Percolation water

Lowland cave system

Figure 2. Relationship of speleothems to external climate-drivers. From Houghton et al., 2001. 22 1.5.2 Physical Weathering and Incision

Many streambeds are at least partially mantled by alluvium. For bedrock erosion to occur, the bedrock must be partially weathered (Howard, 1998). The erosion rate should be proportional to the quantity of sediment in transport when the quantity of bedload is small. However, when the rate of sediment input is high, grains interfere with each other and begin to armour the bed. At which point, the rate of erosion reaches a maximum and presumably goes to zero as the bed becomes completely covered by alluvium (Howard, 1998). In bedrock streams, such as most cave streams, the main concern is the rate of bed scour (Howard et al., 1994).

Abrasion reduces the size of particles and the sharp corners and edges are worn smooth during transport. Transport and depositional processes often results in sorting, which refers to the particle-size distribution (Howard et al., 1994; Monroe and Wicander,

2006). Plucking, abrasion by sediment, solution, and weathering are mechanisms by which erosion can occur. Depending on channel hydraulics, climate, rock type, water chemistry, and sediment type and load the relative importance of these processes can vary

(Howard et al., 1994).

Channel erosion generally involves more than just hydraulic processes such as plucking, abrasion, and solution. The rate of bedrock erosion is affected by weathering, mass wasting, and burial by sediment cover (Howard, 1998). Bedrock streams physically incise by abrading and/or quarrying (Springer, 2002a). Often, the dominant process is abrasion (Howard, 1998). Abrasion is where relatively small moving particles detach more particles and quarrying is the detachment of blocks from river beds and banks

(Springer, 2002a). Short in sedimentary rocks may be an example of 23 quarrying or undermining of resistant beds (Howard, 1998). Quarrying and abrasion are related to aspects of channel flow or hydraulics. The forces required to quarry blocks are reached once stream discharges increase above some critical threshold; abrasion is common atop resistant strata. Both processes are mechanical phenomena driven by gravitational energy (Springer, 2002a). Local exposure of bedrock permitting continued erosion may be provided by migrating bedforms such as dunes, bars, and sediment waves

(Howard, 1998).

Landscape erosion is strongly controlled by the balance between the removal of debris by transport processes and the breakdown of rock into movable material by weathering (Anderson et al., 2002). Physical processes expose bedrock to hydrologic processes and drive soil production. Chemical processes attack the rock vigorously, but this must be viewed as a consequence of the environment produced by physical processes

(Anderson et al., 2002).

The removal, transportation, and deposition of weathering products in almost every terrestrial environment is controlled by (Howard et al., 1994).

Deposited sediments may be separated from the source by thousands of kilometers. The erosion of loose sediment is not equivalent to bedrock incision. The important issues for alluvial channels are quantifying sediment contribution from local erosion and bed sediment transport rate. Bedrock channels occur when a stream is capable of transporting more sediment than it is supplied from upstream and from local slope erosion (Howard et al., 1994). 24 1.5.3 Sediment Transport

Sediments are transported in several ways. The main ways in which transportation occurs is either by chemical or physical erosion. Physical erosion can be further broken down into the many processes by which sediments reach their destination such as by wind, water, or ice movement. The mechanics of sediment transport have been particularly well studied and are described below.

1.5.3.1 Stream Power

Stream power is the kinetic energy available for erosion and transport as described by the reaction:

-1 Ω = γQSe where Ω is stream power per unit length of channel (watts=N m s ), γ

-3 3 -1 is specific weight of the fluid (~9800 N m ), Q is stream (m s ) and Se is energy slope or rate of energy loss per unit length of channel (approximated by channel slope in m/m) (Bagnold, 1966).

However, unit stream power is defined in terms of power per unit area across the channel bed; it quantifies the rate of loss of potential energy as water flows down slope per unit channel width:

ω = (γ Q S)/w = γ D V S = τ V where ω is unit stream power (W/m-2), w is flow width (m), D is mean flow depth (m), V is average velocity (m/s) and τ is shear stress acting tangential to the channel boundary

(N m-2). Unit stream power, as opposed to stream power, may be a particularly useful tool

for evaluating the ability of a river to quarry and abrade (Springer, 2002). A variety of hydrologic factors such as drainage area, channel slope, flow width, season, vegetation, surficial materials and bedrock geology should be taken into account when using these 25 equations. For a given drainage area such as the Buckeye Creek basin, an increase in channel slope or reduction of flow width produces a proportionate increase in unit stream power, the energy available for valley-floor erosion. Accordingly, transport in wider valleys results in lower stream power and promotes sediment storage (Taylor and Kite,

2006).

Stream power is a direct function of discharge and channel slope (Taylor and

Kite, 2006); it represents the actual energy available for sediment transport. Resisting powers represent the critical threshold energy needed to initiate transport in the form of elements that control energy expenditure, such as clast diameter, sediment volume, bed roughness, and woody debris. Energy is expended by transporting sediment and maintaining flow against resisting elements (Taylor and Kite, 2006).

Bedrock-floored channels imply that systems are under capacity with respect to sediment loads and alluvial-dominated channels imply deficits of stream power (Taylor and Kite, 2006). High sediment supply results in alluvium-covered reaches whereas increased transport capacity or reduced supply induces erosion to bedrock (Montgomery et al., 1996; Taylor and Kite, 2006).

As stream power increases (for example, if discharge increases) more energy is available for reworking channel materials, however, in a bedrock stream with no sediment transport, the stream power is expended though frictional processes. In alluvial channels with mobile boundaries some amount of stream power is used for transporting sediment (Gordon, et al., 2004). Therefore, there is an inverse relationship between the volume of sediment stored along a specific reach and unit stream power (Ritter et al.,

2002). 26 From studies of sediment transport mechanics it is known that transport of sediment grains by free-surface flow does not occur until a flow strength threshold is exceeded. Likely, cave channels form where the shear stress generated by surface flow during storms is just large enough to break through bed armour to erode and transport sediment particles (c.f. Tucker and Slingerland, 1997).

1.5.3.2 Shear Stress

Shear stress is maximized at the deepest part of a channel because of its dependence upon channel depth (Springer, 2002a); it signifies the downslope component

of the fluid weight exerted on a particle as motion begins (Ritter et al., 2002) as expressed

in the equation: τ = γ h Se.

Shear stress typically increases with discharge, but is unevenly distributed within

a channel as well as along a channel reach (Kirchner, et al., 1990; Gordon et al., 2004).

When a particle is on the verge of movement, the shear force acting to overturn it is

balanced with the submerged weight of the particle, which holds it in place (Gordon et

al., 2004).

In general, silt and (fine-grained sediment) is transported within the water

column by the supporting action of turbulence and may travel directly from the place of

erosion to points far downstream without intermittent stages of deposition (Ritter et al,

2002). This travels at a velocity slightly lower than that of the water.

Except for short spans of suspension, coarse sediment usually travels as which refers to sediment transported close to or at the channel bottom by bouncing, sliding, or rolling (Ritter et al, 2002). 27 Coarse particles may travel in true suspension, but they are likely to be deposited more quickly and stored semi-permanently as bedload within the channels (Ritter et al,

2002). The duration for which coarse debris remain stationary within a channel varies based on several parameters including the exposure to flow, the interlocking relationships between the particles, the nature of the debris (e.g., its density, shape, and size), and the flow characteristics of the river; with the result that such debris is generally immobile.

As a result of fluctuations in discharge, a single particle may be part of either the suspended load or the bedload at any given time. Generally streams can carry more fine- grained sediment than they actually do with the implication that the concentration of fines is not a function of transporting power, but is a function of supply (Ritter et al., 2002).

During times of low discharge only fine-grained sediment in the water column will be transported (Gordon et al., 2004). On the other hand, coarse sediment is usually available in amounts greater than a stream can carry; therefore its concentration should correlate with the parameters of flow (Ritter et al., 2002).

As stated, for many particles, motion consists of periods of rest alternating with periods of activity (Gordon et al., 2004). This implies that any given particle will be eroded, transported, and deposited many times prior to reaching its destination. A critical energy level must be reached prior to erosion or transportation of channel materials since large amounts of stream energy is expended by overcoming frictional resistance along its bed and banks. Through increasing discharge up to a certain level, particles on the banks and/or bed will begin to erode. Once this level or threshold is reached, sediment transport will begin. Suspended load follows a path dictated by the spiraling movement of the fluid, but bedload tends to roll in the direction of the bed slope. At some point, 28 equilibrium between the amount of sediment and the energy available to carry it is reached. As stream energy decreases deposition begins to occur; for bedload this is when the material settles out of suspension. In motionless water, a particle will be deposited at a rate dependent on its terminal velocity (Gordon et al., 2004).

Karst receive inputs of water as well as suspended sediment and bedload.

Runoff from overlying caprock may flush sediment down vertical shafts, sinking streams or sinkholes and diffuse through overlying soils may transmit soils vertically into the underlying carbonate as described by Bosch and White (2004). At which point these materials are commingled to yield the modern day cave sediment packages.

Lowering of base level results in flow paths being abandoned resulting in higher elevation, dryer, ancestral cave passages. Post-abandonment the sediment deposits will not be exposed to erosive forces that might rework them thus preserving the final episode of deposition (Bosch and White, 2004).

In summary, the same processes which work to erode on the surface also work in caves. The transport of bedload commonly results in abrasion or corrosion of the limestone which helps the cave stream to incise. However, many of the sediments which are into a cave stream end up deposited in that cave. In cases where sediments have already been deposited they can later be eroded and transported again thus removing any armouring which may have formed and in turn promoting incision.

1.6 Purpose

Understanding linkages between past climates and landscape processes is essential to predicting how landscapes will respond to on-going climate changes

(Thompson, 1999; Dyurgerov et al., 2000; Leigh, 2008). Dating and studying how a karst 29 watershed responded to Pleistocene climate fluctuations and periglaciation will improve upon the range of dates already available for Buckeye Creek Cave. The timing of cave stream infilling (aggradation) can be compared to glacial climates recorded in BCC stalagmites (e.g., Granger et al., 2001; Springer et al., 2009). Rates of channel infilling and downcutting (incision) determined from sediments, can then be used to estimate erosion rates on surrounding landscapes at the times of deposition and during particular climate regimes (Palmer, 1987; Tebbens et al., 1999; Reusser et al., 2004; Leigh 2008).

Sediment and stalagmite ages, in combination with the results of an ongoing paleoclimate study being conducted using 18O and δ13C values in BCC stalagmites

(Springer et al., 2008) can be used to determine how rapid climate changes affected

hillslope stabilities in the karstified landscape. As such, the rates at which caves and

landscapes respond to climate changes as well as overall landscape response times can be

determined.

1.7 Research Questions and Hypothesis

The primary research question to be addressed is: Do incision rates calculated using BCC passages tell us about karst processes or long-term incision? In order to address this question three research hypotheses have been proposed.

H1 Aggradation primarily occurs during the transition to glacial periods.

In eastern North America, large temperature decreases during glacial periods are associated with major decreases in precipitation (G.S. Springer, unpublished data), which

lowers sediment transport capacity, but increase sediment supply resulting in aggradation

(Granger et al., 2001; Springer, 2005; Springer et al., 2009). Sediment will accumulate in

stream passages if driving forces are incapable of moving all sediment supplied to a 30 channel and this can be recorded as sediment banks on channel margins, thick deposits beneath streambeds, and terrace-like deposits in abandoned cave passages (Ford and

Williams, 2007). These deposits may result from climate changes where it becomes more arid or when excess sediment is supplied to caves or nearby rivers by tributary streams, glaciers, or landsliding (Vandenberghe, 1995; Springer et al., 2009). If climate becomes so cool that soils freeze and become permafrost, periglaciation may occur and the vegetation that was holding the soil in place will be replaced by sparser plant communities or bare ground. This can release abundant soil into Appalachian streams and rivers and commonly results in aggradation (Eaton et al., 2003).

H2 Incision primarily occurs during the transition to interglacial periods.

On a broad scale, the late Cenozoic transition to glacial-interglacial cycles

beginning 3-4 million years ago may have led to a widespread acceleration of fluvial

incision rates. For example, in eastern North America, fluvial incision rates deduced from

terrace sequences appear to have increased since the late Miocene, perhaps in response to

long-term cooling (Hancock and Kirwan, 2007; Mills, 2000). Caves may preserve

evidence of long-term records of geomorphologic evolution; such as paleoclimate,

erosion, and paleohydrologic conditions. The occurrence of multiple levels in a cave

system are the results of cave streams episodically abandoning stream passages for a new

passage at a lower elevation, commonly at or just below the water table (Palmer, 1987;

1989; Granger et al., 2001; Anthony and Granger, 2004). Such caves are associated with

surface rivers and may record previous base-level positions and long-term records of

stream erosion (Springer et al., 1997; Granger et al., 2001; Stock et al., 2005a,b). 31 Increasing runoff and decreasing or stable sediment yields might result in incision. When the climate begins to warm vegetation growth is initially delayed with the result that runoff and sediment yields are high (Vandenberghe, 1995; Ward et al.,

2005; Leigh, 2008). High erosion rates often occur during periods of increased precipitation (Fuller et al., 2009). As the climate continues to warm and vegetation returns, the sediment load is reduced, possibly resulting once again in incision

(Vandenberghe, 1995; Eaton et al., 2003; Springer et al., 2009). These observations suggest that long periods of stability separate short episodes of instability and stream incision (Vandenberghe, 1995). This might result in rapid, episodic abandonment of cave passages (Johnson and Gomez, 1994; Granger et al., 2001). The variations of stream incision rates through space and time are of great interest, particularly because they reflect the influence of tectonics and climatic change on fluvial processes in the eastern

United States which have increased during the past few million years (Matmon et al.,

2003). An alternative explanation, however, is that incision rates measured over longer intervals of time may inherently be lower than those measured over shorter intervals

(Mills, 2000; Wegmann et al., 2002).

H3 Phenomena inside and outside of the karst watershed are responsible for major

episodes of aggradation and incision by Buckeye Creek.

The geomorphic evolution of Appalachian karst is linked to the history of

Pliocene-Pleistocene glaciations. It is hypothesized that cave stream piracies are a climate

driven response to the lowering of the master stream into which the cave stream flows

(Granger et al., 2001). Incision by the master stream, as a result of climate change,

increases local hydraulic gradients, which encourages the development of steeper or more 32 hydrologically efficient flow paths (Hauer et al., 1997; Tucker and Slingerland, 1997;

Lach and Wyżga, 2001; Brouyère et al., 2004; Fuller at al., 2009).

On the order of multiple glacial/interglacial sequences, river and cave development follow general climatic evolution. At the scale of one glacial-interglacial cycle (~100,000 years), climate forces multiple river responses (Vandenberghe, 1995), one such response may give streams greater ability to erode and transport sediments.

Therefore, if there is no increase in the amount of sediment reaching the stream then flushing of sediment already in the channel would take place. Alternatively, if runoff decreases, but the sediment supply remains the same or increases, then the stream would become transport limited and aggradation would occur (Vandenberghe, 1995; Leigh,

2008). If so, this change should be preserved in Buckeye Creek Cave sediments.

Cave levels in Mammoth Cave, Kentucky were controlled by development of the

Green River valley, which in turn was influenced by the Ohio River (Granger, at al.,

2001). Therefore, the geomorphic evolution of the Ohio River, its , and

Mammoth Cave, is intimately linked to the history of Pliocene-Pleistocene glaciation

(Anthony and Granger, 2007). Inferred incision and aggradation events on the Green

River can be related with important climatic changes and drainage rearrangements in the eastern . The sediment chronology determined from cosmogenic 26Al and

10Be shows that Mammoth Cave’s evolution, and, by extension, the incision history of the

Green River, occurred in step with major climate changes and drainage reorganizations

(Granger, et al., 2001). BCC should show a similar correlation, but its record will be

more closely tied to hillslope processes in the karst watershed because periglaciation

significantly affects landscapes, but is not reported from the Mammoth Cave area. 33 2.0 Study Area

2.1 Geology

Buckeye Creek Cave through which flows Buckeye Creek into Spring Creek is located in Eastern North America in a limestone valley surrounded by clastic rocks.

Greenbrier County, the field area for this study, is located in southeastern West

Virginia (Figure 3). The eastern portion of Greenbrier County lies within the Valley and

Ridge Province and displays a well developed trellis or rectilinear stream drainage. The western portion of Greenbrier County lies within the Appalachian Plateau Province and has a dendritic drainage pattern (Price and Heck, 1939). Greenbrier County lies within a northeastward trending synclinal basin with a moderate dip (Figure 4). The Greenbrier

Group is terminated abruptly to the east by a northeast-trending structure which has been interpreted as an anticline or a normal fault zone and the western edge of the syncline is bounded by the northeast-trending Williamsburg anticline (Jones, 1973). The study area, the Buckeye Creek Basin, is a part of the eastern-most Appalachian Plateau (Dasher and

Boyer, 2000; Springer et al., 2003).

Greenbrier County, particularly north and west of Lewisburg, is underlain by carbonates of the Greenbrier Group (Hocutt et al., 1978). Locally, some folding has resulted in very steeply dipping beds, however, most of the Greenbrier Group is roughly flat-lying or gently folded (Dasher and Boyer, 2000; Springer et al., 2003). The

Greenbrier Group varies in thickness and is locally 247 meters (900 feet) thick. It is composed of eight recognized units: Alderson Limestone, Greenville Shale, Union

Limestone, Pickaway Limestone, Taggard Limestone, Patton Limestone, Sinks Grove 34 Limestone and Hillsdale Limestone (Figure 5) (Price and Heck, 1939; Worthington,

1984).

Of primary interest to this study are the medium to thickly bedded micrites and oolites of the Union and Pickaway Limestones (Springer et al., 2003) and which form extensive cave systems (Hocutt et al., 1978). The Union Limestone is a calcite-pure gray, hard limestone which weathers white, and is often crystalline and containing numerous marine fossils. The Pickaway Limestone is a blue to yellow, dark, hard, and sandy deposit immediately below the Union Limestone. It is shaly at the top and massive at the base; occasionally chert nodules are present in the lower part of this member (Price and

Heck, 1939). The Greenbrier Group is overlain by Mauch Chunk sandstones and shales

(Jones, 1973). Scattered ridges and knobs, which rise above the karst plain, are composed of the Mauch Chunk Group and supply it with clastic sediments (Dasher & Boyer, 1997;

Dasher and Boyer, 2000; Springer et al., 2003). 35

Greenbrier River

Greenbrier River watershed

Figure 3. Location of Greenbrier County and the Greenbrier watershed with a geologic map of Greenbrier County (adapted from http://wvgis.wvu.edu/data/dataset.php). 36

Greenbrier River

Figure 4. Location of study site within the Greenbrier watershed (From Springer et al., 2010), the geologic map of the watershed (adapted from http://wvgis.wvu.edu/data/dataset.php). 37

Figure 5. Generalized stratigraphic section of the Greenbrier Group. Buckeye Creek Cave is formed in the Union and Pickaway Limestones. From Dasher and Balfour, 1994.

38

Figure 6. Location of study area. BC is Buckeye Creek, UE is the upstream entrance and stream sink, DE is the downstream entrance, and SC is the . Short dashes indicate subsurface flow paths that overflow into surface channels during moderate and large . Buckeye Creek Cave is the sole outlet for the topographically enclosed Buckeye Creek Basin. The active stream tier is the easternmost passage. Cave map adapted from Dasher and Balfour, 1994. 39 2.2 Landforms

Drainage boundaries have formed around the study area due to the underlying

geology. In addition, while base level for Buckeye Creek has always been Spring Creek,

the level of Spring Creek itself has changed. These boundaries and changes are discussed

below.

2.2.1 Drainage Boundaries

The Greenbrier River, a tributary of the New River, flows southward with a

drainage area of 3,800 km2 (Springer et al., 2003). The Greenbrier Limestone outcrop

trends southwestward and widens to about 15 km within the study area (Gooch et al.,

1979). Spring Creek and Second Creek, the tributaries of the Greenbrier River, along with the Greenbrier River itself, cross the limestone belt, however, most of the subsidiary

drainage flows through subterranean conduits into the larger streams (Gooch et al., 1979).

Incision by the Greenbrier River and Spring Creek into Paleozoic rocks (Springer et al.,

2003) has resulted in a flow route approximately 61 meters (200 feet) below the karst

plain (Jones, 1973). Both the Appalachian Plateau and Valley and Ridge physiographic

provinces contribute to the Greenbrier River which principally flows along strike atop

sandstones, siltstones, and shales from the Paleozoic. Mean annual precipitation declines

downstream from a high of 152 cm year-1 in the western headwaters to a low of 100 cm

year-1 at the river mouth. The Greenbrier River is a perennial, concave, unregulated

bedrock stream with an overall of 0.001 m m-1 (Springer et al., 2003).

2.2.2 Base Level Position

Base level is defined as the point below which a stream is not capable of eroding

or incising. Although the ultimate base level is sea level, there are temporary base levels 40 such as , but in many fluvial landscapes base level is the bottom of the channel bed in the largest stream in a watershed. However, defining base level is more difficult in karst because base level can effectively be below a surface channel and its bed.

Although incision is accomplished by corrosion, abrasion, and plucking, there are different processes that take place between incision of a valley and incision of a tube below the valley. When a valley incises, the river must cut down by eroding the channel bed. The river can migrate laterally to find the easiest route, but it still must cut through the underlying rock. However, rivers that flow in the subsurface can shift to lower elevations without incising through all of the intervening rock which implies that subsurface incision rates may be higher than surface rates.

A fall in base level in a surface channel is accomplished by erosional processes such as increasing the gradient or through migration. Base level fall in underground streams can be accomplished in the same way. However, since incision of underground streams can take place by leaving rock in place, underground base level fall can happen with minimal rock erosion and subsurface incision rate need not be representative of surface or valley incision rates.

Buckeye Creek, the smaller stream, drains a 14 km2, topographically enclosed

fluviokarst basin near Renick (Figure 6) (Springer, 2002b; Springer et al., 2003). It flows

north and northeast across clastic rocks and crosses the Alderson Limestone and upper

Union Limestone before entering Buckeye Creek Cave. The stream sink is within a large

karst depression and blind valley (Jones, 1973). Buckeye Creek reappears as a large

spring on the banks of Spring Creek, which acts as base level for the cave stream (Dasher

and Balfour, 1993; Springer et al., 2003). 41 The surface drainage and topography of the Buckeye basin have been greatly modified by the flow of ground water (Jones, 1973; Hocutt et al., 1978; Dasher & Boyer,

1997). Sinking streams, extensively interconnected cave systems, sinkhole plains with no surface drainage, and large springs at river level are characteristic features of a karst terrain (Jones, 1973; Hocutt et al., 1978). The appearance of Greenbrier rocks on the surface results in rolling topography with sinking and resurging streams, numerous dolines, and karren. The greatest permeability within the Union and Pickaway

Limestones is along fractures and conduit development tends to favor certain stratigraphic zones (Jones, 1973).

Within the Buckeye Creek Basin, an intermittent stream just above the cave and

Buckeye Creek provide the only surface drainage. During times of low discharge,

Buckeye Creek flows across the limestone as an uninterrupted stream. On the surface, the former course of Buckeye Creek is east of the present cave entrance. Within the relatively flat-lying, massive Union Limestone Buckeye Creek Cave developed along strike joints trending NE (Jones, 1973). Buckeye Creek Cave is approximately 6.4 kilometers (four miles) long. Within the cave, the active stream passages is about 1.8 kilometers (6,000 feet) long with about 15 meters (50 feet) of relief (Dasher and Boyer, 2000) and with passages averaging 8-m high and wide (Figure 7).

2.2.2.1 Spring Creek

Spring Creek borders the basin to the northeast and east and has a drainage area of

245 km2. It drains highlands to the west of Buckeye Creek and flows into the Greenbrier

River. At one time Buckeye Creek followed an easterly course to Spring Creek; however

this route has been abandoned in favor of flowing through the cave. Buckeye Creek has 42 incised at least 31m since this former route was abandoned (Springer, 2002a). Spring

Creek flows through clastics as well as carbonates and is a mixed alluvial-bedrock stream depending on location. Spring Creek, like Buckeye Creek, has a concave profile

(Springer, 2002a).

Spring Creek exhibits characteristics of a karst stream. Infiltration into adjacent alluvium as well as the streambed during periods of dry weather results in long stretches of dry streambed (Dasher and Balfour, 1994). Karst springs return the water to Spring

Creek which will then carry the water until infiltration again takes place. Many caves are present along the sides of Spring Creek. During periods of excess precipitation this infiltration to and from Spring Creek can cause flash flooding. Buckeye Creek responds to changes in the level of Spring Creek which, in turn, is affected by changes in the

Greenbrier River. Thus, as Spring Creek incises or aggrades so does Buckeye Creek. Five levels from incision of Spring Creek can be noted inside of the Buckeye Creek-Rapps

Cave system. The downstream reaches of these levels end against the hillside or in breakdown indicating a progressively lower outlet point. The low gradients of the passages have been interpreted as evidence of long periods of gradual incision within

Spring Creek itself (Dasher and Balfour, 1994), but this need not be true, as will be discussed below. 43

Figure 7. Plan view of Buckeye Creek Cave. The black arrows indicate the flow direction while the red arrow indicates the location of the stalagmite BCC-002. Cross-section Q-q is shown below. Adapted from Springer, 2002.

Figure 8. Cross-section of Buckeye Creek Cave along Q-q showing the four tiers (levels). Starting from the top, TGTL is the oldest flow path which was subsequently abandoned with the formation of Turner Avenue, the second highest level, BRT, and the lowermost l evel are stream passages. From Springer, 2002. 44

Figure 9. Map of Buckeye Creek Cave showing the upstream and downstream entrances and the locations of stalagmites BCC-002 and BCC-006. From Springer et al., 2010.

45 2.2.3 Passage Morphologies

Buckeye Creek Cave contains three abandoned tiers and one current tier; the uppermost tier is the oldest (Figure 8). Passages in Buckeye Creek Cave are either phreatic, epiphreatic, or vadose. Phreatic cave passages are permanently flooded to their ceilings and pressure head drives flow in a manner analogous to pipes (Springer, 2004).

Epiphreatic cave located at the water table, but are only pipe-full during floods. Most of

BCC upstream of the Watergate can be classified as epiphreatic. Vadose cave passages are partially air-filled and governed by the same driving forces as those found in surface streams. This describes the BCC stream passage directly downstream from the Watergate and ending 304m (1000 feet) later. Rapid incision can cause phreatic and epiphreatic passages to be abandoned in favor of narrow vadose canyons (Palmer, 1987; Granger et al., 2001) and aggradation can partially or completely infill caves with sediment (Granger et al., 2001; Springer et al., 2009).

Multiple passage levels, such as found in Buckeye Creek Cave, can be used to interpret cave and landscape evolution (Palmer, 1987). Cave springs along the river banks tend to remain stable when base level is static. Thus, base level passages can grow to relatively large sizes. However, frequent diversions of groundwater to lower elevations can be caused by rapid valley deepening resulting in abandonment of passages at or near river level or entrenchment of cave floors to form canyons. These cause growth of the original water-filled sections to be arrested (Palmer, 1987).

The concave profile and downstream profile of Buckeye Creek has remained constant for the last 1 Ma (Dasher and Balfour, 1994; Springer et al., 2003). This stability implies that channel geometry is adjusted to valley wall interactions and substrate 46 (Springer, 2002a; Springer et al., 2003). A significant lowering of base level during development of the cave is indicated by abandoned passages located more than 30 meters

(100 feet) above the present stream (Jones, 1973).

2.2.4 Sediment Packages and Speleothems

Sediment packages have been deposited throughout Buckeye Creek Cave.

Composed of cobbles, , silt, and clay, the packages are variable in lateral and vertical extent. The sediment that was collected and analyzed for this study was located along the

Berry-Rutzen Trail (BRT), which includes sediments that contained a freshwater snail shell. Throughout much of this level, the lowest inactive passage in the cave, deposits of cobbles can be observed. Two sites were selected for sample collection, the first is herein referred to as Snail Shell Site and the second will be referred to as Stope’s Top, which is equivalent to the cave’s fourth level.

In addition to sediment deposition, throughout Buckeye Creek Cave, stalagmite growth has been and still is taking place (Figure 9). A number of these stalagmites have been dated, as will be discussed below. 47 3.0 Methodology

In working toward completion, five processes were employed. They were surveying, sieving, paleomagnetism, U/Th age dating, and organic carbon.

3.1 Surveying

The original survey of Buckeye Creek Cave was completed by Dasher and

Balfour (1994), however, for this study a more precise survey was required. Distance and inclination were measured using a handheld Leica DISTOTM D3 survey tool. Not only

were passages surveyed, but also the locations of cobble deposits, breakdown blocks,

ceiling height, and speleothem locations. The speleothems that were surveyed were either

surveyed as the location from which they were removed for dating or else where they are

currently located within the cave. Measurements of the ceiling height can be used to

determine the slope of the passage so long as the ceiling is solutional, not collapse.

Breakdown blocks were surveyed because they are stationary features in the cave and can

be used for reference. The locations of the cobbles were surveyed in an attempt to

determine how extensive the deposit was. In addition, this survey proved to be useful for

determining passage elevations. These elevations can be determined by measuring

elevations above the current stream passage. For this purpose, measurements down to the

streams edge were collected.

For surveying, the DISTOTM was placed on a leveled tripod. A reference point

was selected and surrounding measurements were collected of the cobbles, ceiling or

stalagmites. Common reference points included breakdown blocks or marks on the cave

walls from the original survey. Upon moving to the next survey location the distance and

inclination back to the previous reference point was determined and a new reference 48 point was selected. In this fashion the distance and inclination of the relevant passages and deposits was collected (See the Appendix).

3.2 Sieving

At several locations bulk samples were collected. The samples were spaced approximately 20 centimeters apart and collected with the use of a trowel and one gallon plastic bags. The bags were filled approximately half full. Each bag was labeled and taken back to the laboratory for grain size analysis.

Once back in the laboratory the samples were put into aluminum tins and dried in the oven at 60° C for approximately 24 hours. The sieves were all individually weighed and the mass recorded. After the excess moisture had been removed the sediments were placed in a column of sieves. Ranging in size from a 15.875 millimeter (0.6250 inch) sieve to a 0.0063 millimeter (0.0025 inch) sieve. The dried sediment and the column of sieves were then placed in a shaker for approximately 15 minutes. The sediment on each sieve was weighed and the original mass of the sieve was subtracted yielding the mass of the sediment. As the masses were collected they were recorded in a table. After each sediment sample had been weighed a total mass of sediment was obtained and a percent total was derived according to the equation:

(mass of sample ÷ total mass of all samples) × 100 = percent total

The percent finer was then obtained by subtracting each percent total from 100. A table was constructed from these data. For the determination of percent total and percent finer, the mass used was in grams, not kilograms. 49 3.3 Paleomagnetism

The Earth’s magnetic field reverses at irregular intervals during which time the magnetic North and South poles exchange places. Intervals during which the North Pole is near its current location are called ‘normal’; the other intervals are called ‘reversed’

(Sasowsky, 2005; White, 2007). Rocks and sediments can be deposited with magnetic grains parallel to the ambient magnetic field and this is referred to as “primary magnetization” (Sasowsky, 2005; Stock et al., 2005; Hajna et al., 2008). The primary signal points in the direction of magnetic north at the time of deposition (Sasowsky,

2005).

Primary magnetization can be measured to reconstruct the direction and intensity of the Earth’s magnetic field when the sediments were deposited. Using this approach, a chronology of the pole reversals has been constructed (Hajna et al., 2008). The ages of sediments may be determined by measuring magnetic properties of samples and placing them in their proper positions on the magnetostratigraphic timescale (Sasowsky, 2005).

This is done by measuring two magnetic field components: declination and inclination.

Declination is defined as the angle between magnetic north and true north at a particular location. In addition to declination, the magnetic field has an up-down component

(inclination). At the North magnetic pole, the inclination is +90°; at the South Pole it is -

90° and at the equator it is 0°. The primary magnetization is defined by a sample’s inclination and declination (Sasowsky, 2005; Hajna et al., 2008). Five signals are commonly observed: weak and indeterminate, weak and reversed, weak and normal, strong and reversed, or strong and normal (Schmidt, 1984). 50 The Earth’s field periodically shifts and reverses; so primary magnetizations can be affected by a later, opposing field if exposed to it for a long time (Verosub, 1977;

Butler, 2004). This is known as overprinting and results in a secondary magnetic fabric, which is called secondary natural remanent magnetism (Butler, 2004). The remaining original field is called primary natural remanent magnetism (Palmer, 2007). Natural remanent magnetization (NRM) is the overall fabric present in a sample prior to laboratory treatment and is the component sought in paleomagnetic investigations

(Butler, 2004). The geomagnetic field and geological processes during and after sediment deposition affects the NRM (Butler, 2004).

Magnetite is the dominant detrital ferromagnetic mineral in most sedimentary environments. During suspension settling of stream sediments, tiny magnetic mineral grains become oriented with the magnetic field (Sasowsky, 2005; Stock et al., 2005).

When the particles reach the bed, they are buried by additional particles, and the magnetic orientation of the grains becomes locked in (Sasowsky, 2005; Stock et al.,

2005). Still-water environments allow for good orientation of the magnetic particles; this makes clays a desirable material for sampling (Sasowsky, 2005).

The geomagnetic field more efficiently aligns fine particles. Larger particles are not effectively aligned by either depositional or post-depositional processes because they have a lower intensity of magnetization. In addition, they are less likely to move freely within pore spaces in newly deposited sediment (Butler, 2004). Additionally, coarse sediments are permeable and likely to experience chemical changes due to groundwater circulation. For these reasons, clays, silts, or fine sands are preferred in paleomagnetic studies, while larger grain-size sediments are avoided. During compaction and 51 consolidation, fine-grained sediments decrease in water content. Accordingly, there is ample time (perhaps 102–103 yr) for post-depositional alignment to occur (Butler, 2004).

Fine-grained sediments acquire remnant magnetism by preferred alignment of

magnetic grains during deposition or consolidation. Either process may induce an

'inclination error' attributed to compaction effects and grain-size/shape (Løvile, 1987).

After deposition, water-filled voids may be present in the sediments within which

magnetic carriers would be free to rotate. Thus when the water content drops below a

critical value, rotation is halted (Verosub, 1977). The locking-in of detrital remanent

magnetism occurs at this time (Butler, 2004). Depending on sedimentary environment,

estimates of lock-in time range up to 103 years. Fine particles located in interstices are

probably locked in after larger ferromagnetic particles (Butler, 2004).

The NRM is the magnetization first measured in the laboratory. The mechanism

by which the NRM is acquired depends on the mode of formation and subsequent history

of the sample. Various demagnetization procedures are applied to the NRM to distinguish

the primary and secondary magnetization. The two most commonly used techniques are

TD (thermal demagnetization) and AF (alternating field) demagnetization (Hajna et al.,

2008).

The timing of fine sediment deposition in caves can be obtained by

paleomagnetism which provides an important first-order assessment, provided magnetic

reversals of known age can be identified in a stratigraphic sequence (Stock et al., 2005).

Paleomagnetic dating is hindered by a lack of continuous stratigraphy and it involves correlating a local magnetostratigraphic column with the global paleomagnetic record

(Anthony and Granger, 2004; Stock et al., 2005). It is impossible to prove that a 52 magnetostratigraphic section is complete; therefore, when constructing a local magnetostratigraphic column with clastic sediments, it is only possible to determine a minimum age for the deposits (Sasowsky et al., 1995).

The last full reversal, the Matuyama-Brunhes, occurred ca. 0.78Ma; the chronology of magnetic reversals is well established (Cande and Kent, 1995). The presence of magnetically reversed sediments in a cave indicates a minimum age of 0.78

Ma (Stock et al., 2005). Post-depositional mixing and surface weathering effects are reduced in cave sediments because they are biologically sterile environments with a constant temperature regime (Løvile, 1987).

Quaternary sediments can be dated or correlated by using short-lived geomagnetic features (Løvile, 1987). However, the assumptions allowed when using only paleomagnetic data depends on there being unrecognized major hiatuses in the section, and on that assumption, that all reversals at about the same time interval have been previously recorded. Recognition of a reversal may prove useful for correlation between sections even if it cannot be pinned to a particular reversal in the geomagnetic time scale

(Cohen, 2003).

Paleomagnetism suffers from two main limitations as a dating tool. First, it cannot yield absolute ages and is thus is a correlative tool (Stock et al., 2005b). Cave sediment magnetostratigraphy requires extensive sampling of sedimentary sequences extending to the present time. If sedimentation is not continuous, then reversals may not be recorded, leading to misidentification and erroneously young cave age estimates (Stock et al.,

2005b). Second, paleomagnetism is best recorded in fine sediments, the deposition of which may occur after cave development. In addition, these sediments are highly 53 susceptible to remobilization and deposition in cave passages well above base level.

Because cave deposits postdate cave development, deposit ages provide only minimum ages for cave development. Correspondingly, incision rates based on cave ages must be considered maximum rates (Stock et al., 2005b).

Paleomagnetic studies of cave deposits can serve as helpful tools to understand karst evolution and to interpret sediment ages (Hajna et al., 2008). Paleomagnetic reversals in cave sediments provide time markers and thus dates for sediments and the passages in which they occur (White, 2007). The results of paleomagnetic analyses are usually applied to the dating of infilling processes and are also a means of reconstructing climate changes and landscape evolution, especially when combined with other dating methods (Hajna et al., 2008).

For paleomagnetism, duplicate samples and samples from many levels and locations are usually collected. In caves, undisturbed and carefully oriented samples of sediments are obtained by sliding small cubic plastic containers over them after the surrounding material is sliced away (Sasowsky, 2005; Palmer, 2007). Samples are collected without the use of metal instruments to prevent contamination. The samples are then taken to a laboratory and measured in a magnetometer, such as a 3-axis 20G superconducting rock magnetometer. This instrument determines the strength and the direction of the sample (Stock et al., 2005). In the magnetometer, the samples are rotated and their magnetism is measured. The induced field remains oriented along the Earth’s field, but the remanent magnetism rotates with the sample. Differences between the two indicate the orientation of the Earth’s field at the time of sediment deposition (Palmer,

2007). Several steps of magnetic cleaning, or demagnetization, are usually needed to 54 reveal the remanent magnetism (Sasowsky, 2005) because it is easy to miss a reversal where sediments are poorly exposed or have been removed by erosion. Analysis of a sample does not provide a single absolute age, but rather a range of possible ages

(Palmer, 2007). This method is best applied to tiered caves whose total sequence of cave passages and sediments formed over long time spans because there are long time spans between individual reversals (White, 2007).

We can not know what interval a specific sample is from (Sasowsky, 2005; Stock et al., 2005b), but we can make an educated guess using other information, such as U-Th dates of speleothems or landscape position, thus we can narrow the probable age range of the cave sediments and passage (Sasowsky, 2005; Stock et al., 2005). The primary advantage of paleomagnetic dating is that it has a greater range than U-Th disequilibrium dating (Sasowsky, 2005).

3.4 Uranium-thorium

Carbonate speleothems can be dated by exploiting the 238U – 234U – 230Th disequilibrium decay series if they contain ppb-ppm levels of thorium and uranium.

Accurate ages are possible if system has remained closed to post-depositional exchange of uranium and thorium and if the initial concentrations of 230Th are well constrained

(Dorale et al., 2004). Speleothems are formed by meteoric drip water and may be dated

by the U-series method (typically 234U – 230Th). Since drip-type speleothem growth is

episodic, speleothem ages may significantly underestimate cave ages and thereby

overestimate rates of landscape evolution. Secular equilibrium in the U/Th system is typically reached after ~0.4 Ma (Dorale et al., 2004; Stock et al., 2005). Stalagmites can

grow in cave passages anytime after they are drained of groundwater. In fact, speleothem 55 growth is often initiated by climate change rather than water table lowering, (Stock et al.,

2005). Sampling the oldest available speleothems is often difficult because in many cases older deposits are covered by younger deposits.

For many speleothems, initial 230Th concentrations may be trivial (Dorale et al.,

2004). Uranium may be incorporated into speleothem-depositing infiltration waters, but

thorium will not be incorporated due to the geochemistry of these two elements (White,

2004; White, 2007). Uranium is often oxidized to the U6+ state where it appears as the

2+ UO2 ion. The uranyl ion is intrinsically soluble and forms carbonate complexes that further mobilize the element in groundwater (Langmuir, 1978; White, 2004; Palmer,

2007; White, 2007). Thorium is locked into the insoluble Th4+ state and is immobile in

groundwater (Langmuir and Herman, 1980; White, 2004). Thus, any thorium found in a

speleothem has formed by the decay of uranium and is therefore a measure of the time

since the speleothem was deposited (White, 2004; White, 2007; Palmer, 2007).

After collection, a stalagmite can be sectioned along its long axis, and then

sampled at close intervals to yield a complete chronology of the stalagmite’s growth

history (White, 2007). An estimate of the probable range of initial 230Th/232Th values

combined with the measured 232Th/238U ratio reveals whether initial 230Th is significant

(Dorale et al., 2004). If significant, initial 230Th can be constrained by monitoring 232Th and employing isochron techniques (Dorale et al., 2004). The sensitivity of the age error to uncertainties decreases with increasing age, increasing U concentration, and decreasing detrital contamination (Dorale et al., 2004). Closed-system behavior can be judged by stratigraphic ordering of ages, petrographic considerations, and 230Th concordance

(Dorale et al., 2004). 56 Materials younger than 600,000 years and as few as tens of years in age are potentially dateable by the 238U – 234U – 230Th method. By applying these techniques to

appropriate materials, one can obtain accurate and precise ages over the past half million

years (Dorale et al., 2004). Speleothem sub-samples are nearly ideal candidates for 230Th dating (Gascoyne, 1992a). However, three issues with this method are (1) the level of initial 230Th in samples and methods of correcting for initial 230Th, (2) the precision with

which 230Th ages can be determined, and (3) the degree to which diagenesis may affect the accuracy of the ages (Dorale et al., 2004; White, 2004). The number of disintegrations

per unit time (dN/dT), of any radioactive nuclide is equal to Nλ, where N is the number of atoms, T is time, and λ is the decay constant for that nuclide. The half-life of that particular nuclide is equal to (ln 2)/ λ. 234U and 230Th are the longest-lived intermediate

daughters in the uranium thorium decay series. The half-lives of the uranium thorium decay sequence is: 238U → (4.468 x 109y) 234U → (2.453 x 105y) 230Th → (7.569 x 104y)

206Pb (Dorale et al., 2004).

Uranium-thorium dating of speleothems is possible because of the extreme

fractionation removal of Th in ground water. Uranium is soluble as the uranyl ion, and as

various uranyl carbonate complexes, but Th is locked into the Th4+ state with an extremely low solubility. Therefore, surface waters have very low 230Th/238U ratios

(Dorale et al., 2004; White 2004; Palmer, 2007). We can assume that as a speleothem

grows it includes U into its crystal lattice, but not 230Th. If the crystal lattice remains in a

closed system the equations for radioactive productions and decay of 238U, 234U, and 57 230Th govern the geochemical evolution of the system as follows:

230 234  Th   U (m)    1  e 230T    230 1  e (234 230 )T , and  238      U   1000  230  234 

234 234 234T  U (i)   U (m) e .

In these equations, λ denotes the decay constants, [230Th/238U] denotes the 230Th/238U

234 234 238 activity ratio (δ U(m) = ([ U/ U]-1)×1000), and T is the age (Dorale et al., 2004). The

first equation shows that the age can be calculated if the decay constants are known and if

230 238 234 [ Th/ U] and δ U(m) can be measured. The second equation relates the measured

234 234 δ U value to the initial state (δ U(i)) when the system was isolated with no thorium. As

T becomes large, [230Th/238U] approaches unity, and at some point between 400,000 and

800,000 years an age limit is reached for the technique. The exact limit depends on

several factors, including the initial uranium and the precision of isotopic measurements.

Reported uncertainties in T are introduced primarily by uncertainties in the measured

230 238 234 values for [ Th/ U] and δ U(m), but also depending on T itself, because the slope

generally decreases with increasing T, approaching zero near several thousands to several

230 238 234 tens of thousands of years old. The analytical errors in [ Th/ U] and δ U(m) generally follow counting statistics (Dorale et al., 2004).

Uncertainties in chemical blanks, filament blanks, sample weight, spike weight, and spike concentration are typically small when compared to the analytical uncertainty.

However, these uncertainties are also included in the error propagation through equation one. Obviously, accurate values for half-lives are required for accurate age calculations.

Uncertainties in decay constants are typically not propagated through the first equation

(Dorale et al., 2004). 58 For uranium thorium dating by mass spectrometry, the factors that determine a

234 given level of precision include (1) sample age, (2) δ U(i), (3) U concentration, (4) ionization/transmission efficiency, and (5) chemical yield (generally >90%). The number of atoms actually detected versus the number of atoms introduced to the system (a variable number dependent on technique) is referred to as the ionization/transmission efficiency (Dorale et al., 2004). In inorganic calcite work, such as that with speleothems, both sample size and precision are related to the error in age (Ludwig et al., 1992). Since deposition rates of speleothems are often low, small sub-samples are needed. In cases where large sub-samples are used this will integrate large time intervals and introduce age errors if uranium concentrations and growth rates are not constant in that interval. On the other hand, a higher precision result may be provided by large sub-samples because they contain more 230Th atoms. In this, there is an unfortunate tradeoff between spatial

resolution in sub-sampling and precision in the 230Th age. The best scenario is one in

which there is a balance between sampling resolution and analytical precision resulting in

the collection of the most scientifically useful result.

Because speleothem growth depends on a constant supply of drip water, hiatuses

in speleothem growth are often interpreted as an interruption of the water supply. Such an interruption could be the result of either increased aridity at the land surface or of frozen ground and glacial advance. In caves of the northern latitudes, the deposition of speleothems has been punctuated by both advances and retreats of the many Pleistocene glaciations (White, 2004). The mechanisms for the shut-down of speleothem growth in periglacial and/or glacial climates includes blockage of seepage water by perennially 59 frozen ground, blockage by ice cover, and dramatic decreases in biogenic CO2 because of

sparse vegetation and cold temperatures (White, 2004).

When the goal is to accurately date cave development, the best approach for

obtaining speleothem ages is to date several speleothems from each level, if available,

and use the oldest age for calculating an incision rate (Stock et al., 2005). Specific

stalagmite samples from Buckeye Creek Cave were collected prior to this study and

during this study for U-Th dating. After collection, the speleothems are cut in half using a

rock saw and cleaned with deionized water. Approximately 1 gram of calcite is cut from

the basal layer of each sample, detrital grains are removed by hand, and the calcite is

dissolved in 8N HNO3 (Stock et al., 2005). Samples can then be measured by either

thermal ionization mass spectrometry (TIMS) and/or inductively coupled plasma mass spectrometry (ICPMS). The samples are then spiked with either 236U and 229Th, or 233U and 229Th, respectively and run through the spectrometer (Stock et al., 2005). This results in an age for the stalagmite and thus a minimum age for when the passage was drained of water.

3.5 Carbon

13 Plants discriminate to varying degrees against CO2 during photosynthesis (Table

2). Photosynthetic type C3 is more discriminatory than C4 and C3 plant tissues have low

13C values (<-25‰). Natural isotopic differences allow carbon derived from each

photosynthetic type to be traced through the soil organic matter (Boutton et al., 1998).

Alternations of C3 and C4 plants accompany climate changes such as wet/dry cycles and

glacial/interglacial transitions and these changes are evident in δ13C values of organic

matter (Meyers and Lallier-Vergès, 1999). δ13C values reflect the relative contributions of 60 plant species with C3 and C4 photosynthetic pathways to community net primary productivity (Boutton et al., 1998; Zhaoyan et al., 2003).

The natural abundance of C3 and C4 plants is affected by multiple environmental factors including moisture (precipitation), temperature, and atmospheric CO2 (Zhaoyan et

al., 2003); the relative importance of which may vary by ecosystem (Liu et al., 2005).

Changes in δ13C values of modern soils reflect changes in C3/C4 plant ratios. Past

changes in the relative abundance of C3 and C4 plants may provide information for

paleoclimatic conditions (Liu et al., 2005) since the unique δ13C values of C3 and C4

plants are not significantly altered during decomposition (Meyers and Lallier-Vergès,

1999).

Warm humid conditions are often associated with C3 vegetation (trees and

shrubs) while cool arid conditions are often associated with C4 vegetation, such as

grasses (Zhaoyan et al., 2003; McDermott, 2004), many of which are drought tolerant

(Panno et al., 2004). C4 plants have an advantage over C3 plants in that at times of low

pCO2/O2 ratios they are better situated to survive. However, in addition to this, temperature and aridity are also important factors in controlling C3/C4 plant ratios (Liu et al., 2005). If precipitation is sufficiently high enough, even during times of low atmospheric CO2 levels, C3 plants can increase in abundance relative to C4 plants (Liu et

al., 2005). The interchange between C4 and C3 photosynthesis occurs at lower

temperatures when atmospheric CO2 is reduced (Zhaoyan et al., 2003).

In localities where the photosynthetic type has been consistent, the δ13C value of

soil organic carbon in the upper soil profile is similar to that of the plant community

(Boutton et al., 1998). However, changes in the relative proportions of C3 and C4 plants 61 can be recognized as changes in δ13C. Immediately following a vegetation change this

isotopic difference will be largest, but will decrease over time as the carbon from the previous plant community decays (Boutton et al., 1998). C3 plants have δ13C values

ranging from -20‰ to -32‰, whereas C4 plants range from -9‰ to -17‰ (Zhaoyan et

al., 2003; Liu et al., 2005). A site-specific chronology of relative productivity of the

different photosynthetic types can be developed by measuring the δ13C at different depths

in a soil profile. Because of the differences between C3 and C4 plants, the isotopic

composition of soil organic matter can be readily used to differentiate input from the two

types of plants (Liu et al., 2005). The proportion of C4 to C3 vegetation growing

proximal to sinkholes and swallets at the time of redeposition is represented as the δ13C

values of the organic matter (Panno et al., 2004). Since the δ13C in paleosols can persist

without diagenetic alteration for more than 5,000,000 years, this method may also be

applied to cave sediments (Boutton et al., 1998).

Within Buckeye Creek Cave, sediment profiles were measured at two centimeter

increments. At each increment a small metal knife was used to scrape at least two grams

of sediment into plastic bags which were then labeled. Once back in the laboratory the

samples were placed in small aluminum dishes and heated in the oven at 60° for over 24

hours. Once all the excess water has been driven off the samples were ground up using a

mortar and pestle. Approximately one gram was then measured and placed in a small,

labeled plastic bag. The sediments were then sent to Dr. Harry Rowe at the University of

Texas Arlington for analysis according to the method described by Zhaoyan et al (2003).

This method is described herein: once received, the samples are reacted with 1 mol/L

HCl for carbonate mineral removal. They are then washed with distilled water to 62 Table 2. Representative isotopic data from various organic matter (Mackay et al., 2005; Boutton,

1991; 1996).

Organic Matter δ13C

C4 vegetation: Less humid, maize -17 to -9

C3 vegetation: Temperate forests -32 to -20

River Plankton -30 to -25

Lake Plankton -42 to -26

Macrophytes -30 to -12

Peat -27

Lake Shore (common) -27

63 neutrality, the HCl resistant residues are concentrated using a centrifuge, and dried at

80°C. One gram of residual sediment is then mixed with five grams of CuO in a glass vessel which is heated at 550°C for 24 hours. The CO2 is converted from the organic

matter and is collected by liquid nitrogen. The 13C/12C is measured in the MAT-252 mass

spectrometer. By use of this method, the uncertainty (1σ) is less than 0.2‰ (Zhaoyan et al., 2003). Most isotopic compositions are expressed in terms of parts per thousand differences from a standard:

δ13C = [((13C/12C sample)/(13C/12C standard)) – 1] ×103 (Peterson and Fry, 1987;

Zhaoyan et al., 2003). The standard reference material for δ13C is carbon in the PeeDee

limestone (Peterson and Fry, 1987). The δ13C value is a measurement of the amount of

light and heavy isotopes in a sample. An increase in δ13C denotes an increase in the

quantity of heavy isotope components. Conversely, decreases in δ13C values denote a

decrease in heavy isotopic content (Peterson and Fry, 1987). 64 4.0 Results

The results of this study in combination with previous work conducted in

Buckeye Creek Cave are presented in Table 3. This table and the information contained within will be explained in detail in the following sections.

Table 3. Table summarizing incision rates calculated for this study. Event Timing Elevation above Incision rate Evidence present stream (m) (m/Ma) Formation of To- >773000 26.8, <35 Paleomag Good-To-Last/ 29.9 <39 Stope’s Top Abandonment of >773000 26.8 <35 Paleomag To-Good-To- Last/ Stope’s Top Formation of >308000 21, 24.4 <68, <71, <45 BCC-011, Turner Avenue >540000, <50 BCC-022, >489000 BCC-025 Formation of >540000, 24.4 <45 BCC-022, Prism Canyon >489000 <50 BCC-025 Abandonment of >298000 21 <71 BCC-012 Turner Avenue/ Prism Canyon Formation of >539000 16.3 <30, <27 BCC-027 Berry-Rutzen Trail/McClungs Avenue Infilling of >294000 18.3 <62 BCC-013 Berry-Rutzen Trail/ McClungs Ave. Abandonment of >294000 18.3 <62 BCC-013, Berry-Rutzen >539000 16.3, 14.6 <30, <27 BCC-027 Trail/ McClungs Avenue Formation of <539000 BCC-027 Buckeye Mainstream

65 4.1 Elevation Data

An in-cave survey was completed within BCC. Elevations range from zero to

35m above the zero datum (current stream level). The purpose of this survey was to expand on the existing survey and to determine elevations used in the incision rate calculations. The oldest passages, Too-Good-To-Last (TGTL) and Stope’s Top have the highest elevations varying from 26.8 upward. The implication is that water, even in phreatic caves will flow down hill and as a result the stream path is visible on the ceiling of the cave.

After Too-Good-To-Last and Stope’s Top were abandoned Turner Avenue and

Prism Canyon formed. The elevations of these passages vary between 21 and 24.4 meters above the current stream passage. With the formation of the Berry-Rutzen Trail and

McClungs Avenue the stream dropped to an elevation between 16.3 and 18.3 meters above the current stream passage. Finally, the last incision event occurred and the stream cut down to its current elevation of zero.

4.2 Age Data

The ages used in this study were either collected from stalagmites or are inferred from paleomagnetic results. Stopes’s Top and TGTL have been dated using paleomagnetism to reveal a reversed signature indicating the ages of these passages to be over 773,000 years old.

Prior to this study several stalagmites were collected from passages along the level of the Berry-Rutzen Trail and McClung’s Avenue to be dated. Only one stalagmite was collected during the course of this study. BCC-027 was collected in an area where the Berry-Rutzen Trail and Buckeye Creek intersect. The stalagmite was approximately 9 66 centimeters high and 5 centimeters wide at the base. It was no longer growing and its exterior was severely degraded. It was, however, datable and returned an age greater then

539,000 years.

4.3 Incision Rates

Incision rates were determined by using ages and elevations. Because Stopes’s

Top and TGTL are 26.8 and 29.9m above the current stream and have minimum ages of approximately 773,000 years old, their maximum respective incision rates are 35 and 39 meters per million years. It is possible that Spring Creek was above the cave during the formation of these passages. This is indicated by the shape of the passages in TGTL i.e. the roundness of the passages. Thus, the passages were phreatic at the time of formation.

After formation, but before abandonment the passages became vadose and began depositing sediments.

The Berry-Rutzen Trail (BRT) and McClung’s Avenue are at an elevation of

16.3m and 18.3m above the current streambed and sediments in BRT are at least 539,000 years old. This yields incision rates of less than 30m/Ma, <27m/Ma if the sediment thickness is subtracted from the height of the BRT. If the age is calculated from the younger stalagmite, BCC-013, then the incision rate is <62m/Ma. The sediments are at the top of an abandoned phreatic loop (a situation where due to pressure, water actually flows up gradient.), but appear to have been deposited during a major aggradation event.

4.4 Stope’s Top Sediments

At Stope’s Top, three fining upwards units were observed, they consist of pebbles and fine sand bases fining up into fine sands and silts in alternating color layers of orange and grey. The first fining upwards unit consists of pebbles & sands fining up into silts. 67 This fining upwards unit extends 170 centimeters down from the ceiling. The next fining up unit consists of sub-rounded pebbles of various sizes ranging from 2 millimeters to several centimeters; this unit is roughly 22 centimeters high. In the third fining upward unit, there are 1-3 centimeter high occasional lenses of sand & small pebbles; this unit is

123 centimeters thick. In all cases, the course units pinch out laterally and the sediments are highly oxidized.

4.5 Snail Shell Site

The snail shell site is located at the intersection of a passage that runs down into

Buckeye Creek and another that leads to more of the Berry-Rutzen Trail. Four units are distinguishable at this site based on composition and grain size. Unit one is locally 22 centimeters thick. Although oxidized on the surface, the sediment behind the oxidation is thinly laminated silt with a gradual bottom and possible rare burrows and ceiling chips present near the top of the deposit, which is also the ceiling of the cave. There is no evidence of bioturbation, ripples, or cross-bedding and the laminations are well preserved and flat-lying.

Unit two extends from 22 to 63 centimeters deep and is composed of thinly laminated silt and sand with a gradual top and unobserved bottom covered by slump.

Within unit two there are eight sequences of fine sand units that fine upwards into silts.

There is no evidence of burrows, ceiling chips, bioturbation, ripples, or cross-bedding.

Overall, unit two is a fining upwards sequence as more sand is present near the bottom.

Unit three is 74 centimeters thick and is composed of massive silt. The bottom contact is abrupt but the top contact is not observed. There are scattered ceiling chips and soft sediment deformation. 68 Unit four is a 173-centimeter thick cobble fill (Figure 10). The cobbles are oxidized and range in size from 1 to 19 centimeters with an average of 8.9 centimeters.

The tightly compacted, rounded clasts are fairly well weathered and can be broken apart with a rock hammer fairly easily. The top and bottom of unit four is abrupt and the matrix is silty or sandy. Within the unit there are several fining upwards units. The cobbles are imbricated and record a flow direction similar to the modern stream. Laterally, the unit grades into smaller cobbles.

Below unit four is cohesive silt with some clay, some small laminations are present. Samples were not collected from this unit due to sampling difficulty, possible contamination, lacking bottom contact and evidence of slumping indicated that this unit may not be reliable.

4.6 Sieving Results

Twenty-four bulk samples of sediments were collected from within Buckeye

Creek Cave. The sediments have been described in the two previous sections and are

th presented in Table 4. For 10 of the sediments the 16 percentile (d16) is finer than coarse sand, which has a grain size of 1 - ½ mm. The median grain size (d50) for 14 of the

th sediments and the 84 percentile (d84) of 12 of the sediments are finer than coarse sand.

th st d50 is based on a percentile where the largest class is 100 and the smallest is 1 .

Therefore, for d50 half the grains are larger and half are smaller. The mean for d50 is

1.24mm, the median is 0.58mm, the standard error is 0.60, and the 95% confidence

interval is 1.27. The first sample collected contained several large cobbles. Removing these cobbles from the calculations yields a mean of 0.66mm, a median of 0.57mm, a 69 standard error of 0.16, and a 95% confidence interval of 0.34. These values more accurately describe the majority of the sediments.

The maximum particle size that can be carried at a given flow is defined as flow competence (Gordon et al., 2004). For many sediment deposits in BCC the largest grains are less than 2mm, which means that according to the Hjulstrom diagram (Figure 11) they require a minimum velocity of 30cm/sec to be eroded and a mean velocity of

11cm/sec would be required to keep the sediments from being deposited. However, for unit four of the snail shell site a much greater velocity (~200cm/sec) would be required to erode the cobbles and a much higher velocity would also be required to continue transporting the sediments. It is unclear why such large cobbles are not being transported by the modern cave stream.

4.7 Wolman Count

While collecting sediments from the cave a Wolman count was conducted for the cobbles located in unit four of the snail shell site. The results are presented in Table 5.

The cobbles were measured along their a- and b-axes. The mean for the cobbles along the a-axis is 5.61cm while the mean for the b-axis in 3.83cm. The median is 5.5cm and 4cm with a standard error of 0.42 and 0.30 and a 95% confidence level of 0.85 and 0.60, respectively. The smallest a-axis is 1.5cm while the largest a-axis is 10cm. The smallest b-axis is 1cm while the largest b-axis is 6.5cm.

70

Figure 10. Massive cobble fill atop of laminated silts, sands, and gravels in the upper levels of Buckeye Creek Cave. The gravels are >534,000 years old and record a major influx of sediment into the cave, perhaps associated with a change from a dry to wet climate. The photo was digitally enhanced to emphasis the contact between the coarse (upper) and fine (lower) sediments.

71

Figure 11. Hjulstrom curves velocities required for erosion, transportation, and deposition of sediments (Gordon et al., 2004). 72 Table 4. Summary of the sieving results showing the 16th, 50th, and 84th percentiles.

Sample d16 (mm) d50 (mm) d84 (mm) 3/1/09 HF_01 1.300 11.125 3/1/09 HF_02 0.600 1.700 3/1/09 HF_03 1.600 0.400 1.600 Stope's Top HF_09_001, 189-186cm 0.130 0.565 7.000 Stope's Top HF_09_002, 174-167cm 0.170 2.500 8.000 Stope's Top HF_09_003, 152cm 0.160 0.255 0.970 Stope's Top HF_09_004, 281cm 0.190 0.880 3.500 Stope's Top HF_09_005, 265cm 0.353 0.800 1.410 Stope's Top HF_09_006, 243cm 0.460 1.100 1.900 Stope's Top HF_09_007, 225cm 0.275 0.600 1.300 Stope's Top HF_09_008, 205cm 0.200 0.690 1.800 Stope's Top HF_09_009, 305cm 0.310 1.900 Stope's Top HF_09_010, 136cm 0.125 0.195 Unit 3, 30cm, HF_110709_64 0.179 0.395 Unit 3, 40cm, HF_110709_65 0.190 0.390 Unit 3, 50cm, HF_110709_66 0.124 0.353 Unit 3, 60cm, HF_110709_67 0.124 0.195 0.355 Unit 3, 70cm, HF_110709_68 0.150 0.353 average =0.439 1.243 1.534 median =0.238 0.583 0.395

73 Table 5. Wolman count performed on the sandstone cobbles at the Snail Shell Site. Approximately 40% of the cobbles are really well rounded.

Snail Shell Site # a (cm) b (cm) 1 7.50 4.50 2 4.00 1.50 3 5.50 4.00 4 2.50 1.75 5 5.50 4.00 6 3.00 1.50 7 7.00 6.50 8 4.00 4.00 9 6.50 5.50 10 4.00 2.00 11 10.00 6.00 12 6.00 5.00 13 4.50 2.50 14 5.00 4.00 15 6.50 3.50 16 9.00 6.50 17 2.25 1.75 18 2.50 1.50 19 9.00 5.50 20 9.25 5.50 21 4.25 3.50 22 5.00 4.50 23 7.25 4.50 24 9.00 6.25 25 7.50 4.00 26 5.00 4.75 27 6.25 4.00 28 5.25 2.75 29 7.00 4.50 30 2.25 2.00 31 1.50 1.00

74 5.0 Discussion

5.1 Comparison of incision rates

On a broad scale, the late Cenozoic transition to glacial-interglacial cycles beginning 3-4 million years ago may have led to a widespread acceleration of fluvial incision rates. For example, in eastern North America, fluvial incision rates appear to have increased since the late Miocene, as deduced from terrace sequences and perhaps in response to long-term cooling (Mills, 2000; Hancock and Kirwan, 2007). The changes in stream incision rates through space and time are of great interest, particularly because they reflect the influence of tectonics and climatic change on fluvial processes in the eastern United States (Matmon et al., 2003). An alternative explanation, however, is that incision rates measured over longer time intervals may be inherently lower than those measured over shorter intervals (Mills, 2000; Wegmann et al., 2002). This is because higher magnitude events have a longer recurrence interval therefore; rates measured over longer time spans include longer periods of inactivity. This results in a slower rate.

However, the increase in incision rates may instead be attributed to cooler climatic regimes where due to an accumulation of ice on the continents, sea level, the ultimate base level, dropped causing incision (Mills, 2000).

Using cosmogenic isotopes in quartz grains in cave sediments, Granger et al.

(1997) were able to determine minimum ages of caves along the New River in Virginia.

The authors assumed the New River was at the elevation of the individual caves when the sediments were deposited and sediment emplacement times were used to calculate incision rates where the depth of incision was assumed equal to the vertical distances between individual caves and present New River channel bed. Samples were collected 75 from five caves of varying elevations above the modern stream level. The sediment emplacement times varied from 0.29 to 1.47 Ma for caves 12 and 29m above the modern stream. Regression of the vertical offsets by sediment ages yielded an average incision rate of 27 ± 4.5m/Ma. Four of the caves were clustered together while the fifth was located ~10 km to the southeast. The average incision rates for the first four is 30 ±

5.5m/Ma while the incision rate for the fifth cave is 19 ± 3.2m/Ma (Granger et al., 1997).

Previous attempts to measure the New River’s incision rate yielded much faster values,

40m/Ma (Houser, 1981) and 55m/Ma to 100m/Ma (Bartholomew and Mills, 1991).

It is, however, noted that a single river downcutting rate deviates from some of the points by 1.5 standard errors implying that either there is an underestimation of uncertainties or that a single regression line does not explain the data. One such uncertainty is the assumption that the sediment remained unburied prior to cave deposition. Another is the assumption of a constant river downcutting rate, which is contradicted by a terrace study. Using terrace locations and dates determined by cosmogenic radionuclides along the New River, additional incision rates were determined by Ward et al. (2005). Rates of 10Be production can be affected by the density of the

terrace material, topographic shielding, changes in solar output, atmospheric density,

geomagnetic field configurations, as well as shielding by snow, water, loess, and

vegetation. Due to this, the authors assumed minor shielding from topography, snow,

water, loess, and vegetation. They also had to assume a single production rate when a

range of rates is available. In several cases they made corrections to the exposure ages

based on inheritance from erosion and transport of sediment, physical pedogenesis, soil

mixing due to bioturbation, or surface erosion. The authors had to determine which 76 samples needed correction, and then they had to determine how to correct the samples

(Ward et al., 2005).

After collection and analysis of their samples the authors combined their results with those of other studies. The rates determined for the different terraces were between

21 and 80m/Ma with an average of 43m/Ma. However, incision rates have reached as least 100m/Ma during short time periods. Nonetheless, the average rate is consistent with other average New River incision rates ranging from 20-40m/Ma and comparable to the rate of 27m/Ma calculated from cave sediments (Ward et al., 2005).

Springer et al (1997) used magnetostratigraphy of cave sediments to calculate incision rates for the Cheat River Canyon in West Virginia. By collecting and analyzing sediment samples up to an elevation of 51m above river level it was determined that a magnetic epoch boundary was located between 44.5 and 48.8m in elevation. The sediments below 44.5m were normal while the sediments above 48.8m were reversed.

The gap between the two elevations and the missing boundary led the authors to assume that the boundary was the Bruhnes-Matuyama contact as opposed to an older boundary.

Thus, they were able to determine an incision rate between 16 and 38 m/Ma which is comparable to rates determined for Mammoth Cave (33m/Ma). However, the Green

River has a much lower gradient than the Cheat River (Springer et al., 1997).

The authors assumed that all magnetostratigraphic samples are representative of base level, and that the gradient of the river had been constant. In addition, they also assumed that cave sediments behave like terraces and that tiered passages within multi- level caves represent base level stands. Thus the sediments can be used to infer base level. However, this assumption is not valid for vadose passages without additional 77 evidence. Thus, not all cave sediments are equally representative of base level. For example, phreatic sediments can be deposited 20m below base level while vadose sediments can be deposited as high as 20m above base level. Therefore the authors implemented the use of error bars to calculate incision rates which produced a range. This range is between 56 and 63 m/Ma with an average of 59 m/Ma for the past 0.788Ma

(Springer et al., 1997).

The authors then compared their rates to other reported rates; 20m above the

Cheat River are terrace sediments that were analyzed and yielded an incision rate of

<50m/Ma. Downstream of the junction of the Cheat River with the Monongahela River terraces were traced and dated. Incision rates implied from this are <11m/Ma for pre-

Illinoian terraces, 21-50m/Ma for Illinoian terraces, and 20-24m/Ma for early

Wisconsinian terraces. Additionally, at the beginning of the Monongahela River sediments analyzed produced an incision rate of <60m/Ma. Cave sediments provided incision rates of 31-58m/Ma for the Greenbrier river which were then refined to 46m/Ma.

Therefore, a Cheat River incision rate of between 56 and 63m/Ma is reasonable (Springer et al., 1997).

The East Fork Obey River is located in a reentrant part of the Western

Cumberland Plateau. In caves along the Obey River Sasowsky et al (1995) collected samples from 49 sites for paleomagnetic analysis. The sediments up to 49m above river level are all normal and indicate deposition during the current Brunhes chron. From 49 to

60m above current stream level there are normal, reversed, and indeterminate sediments.

This represents the end of the Jaramillo event giving the sediments a minimum age of

0.91Ma. However, since it is impossible to prove a section is complete during 78 paleomagnetic reconstructions there may be a missed boundary. Yet, because of the narrowness of the gorge and the high stream gradient the authors assume no missing boundaries (Sasowsky et al., 1995).

The time-averaged incision rate of the Obey River was thus determined to be

60m/Ma. Since a minimum cave age was used the incision rate is a maximum. The authors assumed a constant incision rate even though there are many factors that may cause an interruption during incision (Sasowsky et al., 1995).

Anthony and Granger (2007) collected samples from twelve multilevel caves along tributaries of the Cumberland River. Four of those caves were along the East Fork

Obey River, the rest of the caves were along four different tributaries. Using cosmogenic isotopes, ages of quartz grains in cave sediments were determined. In addition to assuming a single exposure history of the grains, the authors also assumed that the caves developed in step with elevation changes of regional rivers and that the cave sediments collected represent the last time the passage was at the local water table (Anthony and

Granger, 2007).

From the four caves along the Obey River eight samples were collected. The elevations ranged from 54 to 0m above modern river level (amrl). Burial ages ranged from 1.80 ± 0.31 to 0.02 ± 0.13Ma. Thus the resulting incision rates presented a range.

Ignoring the 0m/Ma calculated by the lowest cave passage the other seven incision rates for the Obey River were between ≤49 and ≤16m/Ma (Table 6). Of the other caves, incision rates ranged from ≤54m/Ma for a passage at 48m to ≤16m/Ma for a passage at

91m amrl (Anthony and Granger, 2007). Thus, the average rate of incision for the East

Fork Obey River would be less than the 60m/Ma as determined by Sasowsky et al (1995). 79 Table 6. Incision rates as determined from age and elevation data from Anthony and Granger (2007).

Elevation Incision rate Age (Ma) amrl (m) (m/Ma) 54 1.23 ≤44

52 1.64 ≤32

48 1.45 ≤33

42 0.85 ≤49

40 1.80 ≤22

28 0.86 ≤33

13 0.83 ≤16

80 Using cosmogenic radionuclide dating of 59 samples from the Susquehanna and

Potomac Rivers, Reusser et al (2004) determined the rate and timing of bedrock incision.

The authors sampled fluvially eroded bedrock terraces. The bedrock preserved fluvial forms allowing the authors to assume that little erosion has taken place since abandonment. Additionally, the authors assume that the ages were not affected by floodwater, snow, vegetation, soil, or topographic features thus the ages obtained were modeled as terrace abandonment ages. From >100 to 32ka the Susquehanna River was incising at a rate of <200m/Ma, then from 32 to 16ka it was incising at a rate of

~500m/Ma. The Potomac River essentially mirrored this. However, further upstream in the Potomac River between 37 and 13ka the incision rate was ~800m/Ma while further downstream the Potomac River was incising at a rate of 500m/Ma from 27 to 8ka. The authors determined these rates were related not only to glacial meltwater, but also to increases in storminess, fore-bulge induced uplift and sediment availability (Reusser et al,

2004).

Whether by paleomagnetism of cave sediments or by cosmogenic isotope analysis many studies pertaining to age estimates and incision rates have been conducted in areas similar to the Buckeye Creek Basin. Incision rates, such as those in the New River, East

Fork Obey River, and Cheat River Canyon are similar to incision rates found for the passages in BCC. Incision rates for the Potomac and Susquehanna Rivers are very different from those determined for BCC. However, even within a single study area rates may vary based on how the ages were calculated, either by cave sediments or terrace exposure. In addition, since the interval of time used for determining an incision rate is also a factor, studies which describe a series of incision events may be more reliable than 81 those which present only an average. As such, studies that present a range of rates are more easily compared with results obtained for BCC.

Since some of the BCC incision rates have been determined by paleomagnetism these dates likely take a larger time frame into account. The rates determined by U/Th take less time into account and are likely more affected by karst processes. Although, in some cases incision may be largely the effect of a few catastrophic events, in the Buckeye

Creek Basin incision is also affected by karst processes. As such, the rates determined by paleomagnetic dating of sediments may reflect more accurate rates because they contain a smoothing of the karst processes. Thus, the rates for To-Good-To-Last and Stope’s Top, which are lower than the rates for several of the younger passages, may be more representative of the rates overall. However, the lowest rate is for the Berry-Rutzen Trail

(BRT), which is much younger than To-Good-To Last.

5.2 Spring Creek

Although the active stream passage in Buckeye Creek is graded to the Spring

Creek surface channel (Springer et al., 2003), its relationship to the Spring Creek surface channel has probably varied through time and this raises questions about what incision rates calculated for inactive passages in Buckeye Creek Cave tell us. For instance, the relationship between the BRT passage and surface Spring Creek ≥539,000 years ago is of particular interest. An incision rate of ≤30m/Ma is calculated if we assume BRT was graded to surface Spring Creek. However, in this case the cave had incised below BRT and later filled with sediment before incising back through the sediment. Therefore, the calculated incision rate overestimates the actual bedrock incision rate, which must be lower than that calculated using the fill sediments. Subtracting the known thickness of 82 sediments from the elevation used to calculate the incision rate, I obtained a rate of

≤27m/Ma. The bedrock incision rate becomes even less certain if we consider the possibility that BRT was graded to an underground Spring Creek (Figure 12).

Upstream of Spencer Cave, the BCC spring, Spring Creek is incising cavernous limestones of the Union Formation and flows underground, except after heavy precipitation when water overflows swallet holes supplying water to the caves and flows through the surface channel. Local subsurface tributaries contribute to the underground base level through vadose and phreatic passages. One such tributary is the Boar Hole –

Portal Cave System, which contributes to the underground Spring Creek where it is ~34.5 meters below the surface channel: the elevation of the underground Spring Creek is 569m

(1,867 feet) amsl and the surface elevation above that point is ~604m (1,980 feet) amsl

(William Balfour, personal communication). However, the Union Formation is underlain by aquicludes that slow or stop underground incision. As such, the surface channel may gradually “catch up” to the subsurface Spring Creek. This would result in very misleading incision rates.

Buckeye Creek last flowed through the Berry-Rutzen Trail over 539,000 years ago and this age yields an incision rate of ≤27m/Ma. This rate is the lowest of all the calculated rates for Buckeye Creek. The low rate can be explained if Spring Creek was flowing through an underground conduit and its elevation relative to base level did not change for a long period of time. The subsurface conduit did not incise because it was likely perched in the Pickaway Limestone or atop the Taggard Shale. Meanwhile, surface

Spring Creek was a losing stream that only carried water during times of overflow much like what is occurring upstream at the Boar Hole – Portal Cave System. During overflow 83 periods the amount of water flowing through the surface stream would have been large resulting in short periods of stream erosion by corrosion, plucking, and abrasion. Over time, the surface stream of Spring Creek incised into the subsurface conduit and the rate is independent of the time required for subsurface base level changes. If so, incision rates calculated using BCC passages tell us about karst processes and not long-term stream incision per se.

However, in light of this finding, it is now necessary to reevaluate the incision rates reported by the authors mentioned above. Incision rates obtained using cave sediments may now be misleading. Since no author has taken into account the possibility of a subsurface base level it is now possible that the previously obtained incision rates are too high and that the actual incision rates with a subsurface base level are much lower.

Thus, since incision rates in a karst terrain reflect karst processes and not long-term incision, rates previously obtained are flawed and future rates must now take this into consideration.

84

Figure 12. Cartoon showing subsurface vs. surface channels of Spring Creek 539,000 years ago.

85 6.0 Conclusions

Evidence within Buckeye Creek Cave preserves a record of incision and aggradation that began over 773,000 years. Incision/aggradation cycles can be linked to surface processes and passage ages can be used to determine incision rates. Incision rates calculated by U/Th or paleomagnetism dating of speleothems and cave sediment varied between ≤27 and ≤71 meters/million years. These rates are consistent with other reported

Appalachian rates. Theoretically, incision by Spring Creek can be inferred using the

Buckeye Creek Cave data. However, the low incision rate of ≤27m/Ma for the Berry-

Rutzen Trail can be explained if Buckeye Creek was adjusted to a subsurface Spring

Creek as base level as opposed to a surface channel and valley. The Berry-Rutzen Trail contains thick packages of sediment. These sediments accumulated in an aggradational event as the result of climatic change. Thus, Buckeye Creek not only had to incise and form the cave passage, but it also had to incise back through all of the sediment which had been deposited. This explains why comparatively little incision has occurred since the original passage floor was excavated. Thus, incision rates calculated using BCC passages tell us about karst processes and not long-term stream incision.

A subsurface base level is one in which the stream is flowing below ground, perhaps underneath a surface channel. However, the surface stream would only carry overflow from the underground stream. The presence of a subsurface base level has not been previously taken into account in incision studies conducted in other Appalachian karst regions. However, the situation at Buckeye Creek Cave is not unique. Thus, it may now be necessary for future and past researchers to re-evaluate incision rates in karst regions as they may not represent long-term rates as previously thought. 86 7.0 References

Anderson, S.P., Dietrich, W.E., Brimhall, Jr., G.H., Weathering profiles, mass-balance

analysis, and rates of solute loss: Linkages between weathering and erosion in a

small, steep catchment, Geological Society of America Bulletin 114, 1143-1158

(2002).

Anthony, D.M., and Granger, D.E., A late Tertiary origin for multilevel caves along the

western escarpment of the Cumberland Plateau, Tennessee and Kentucky,

Established by Cosmogenic 26Al and 10Be, Journal of Cave and Karst Studies

66(2), 46-55 (2004).

Anthony, D.M., and Granger, D.E., A new chronology for the age of Appalachian

erosional surfaces determined by cosmogenic nuclides in cave sediments, Earth

Surface Processes and Landforms, 32, 874-887 (2007).

Bagnold, R.A., An approach to the sediment discharge problem from general physics.

U.S. Geological Survey Professional Paper 422 I (1966).

Baker, V.R., Paleoflood hydrology and extraordinary flood events, Journal of Hydrology,

96, 76-99 (1987).

Bartholomew, M.J., and Mills, H.H., Old courses of the New River: Its late Cenozoic

migration and bedrock control inferred from high-level stream gravels,

southwestern Virginia. Geological Society of America Bulletin 103, 73-81 (1991).

Bender, M., Sowers, T., Dickson, M.-L., Orchardo, J., Grootes, P., Mayewski, P.A.,

Meese, D.A., Climate correlations between Greenland and during the

past 100,000 years, Nature, 372, 663-666 (1994). 87 Bland, W., and Rolls, D., Weatheirng: an introduction to the scientific principles. A.

Hodder Arnold Publication, London and New York, 288pp, 1998.

Bosch, R.F., and White, W.B., Lithofacies and transport of clastic sediment in karstic

aquifers, in: Sasowsky, I.D., and Mylropie, J. (Eds.), Studies of Cave Sediments,

Kluwer Academic, New York, pp1-22 (2004).

Boutton, T.W., Archer, S.R., Midwood, A.J., Zitzer, S.F., Bol, R., δ13C values of soil

organic carbon and their use in documenting vegetation change in a subtropical

savanna ecosystem, Geoderma 82, 5-41 (1998).

Butler, R.F., Paleomagnetism: Magnetic domains to geologic terranes. University of

Portland, Portland, 248 pp. 2004. Baker, R.G., Bettis III, E.A., Denniston, R.F.,

Gonzales, L.A., Plant remains, alluvial chronology, and cave speleothem isotopes

indicate Holocene climatic changes at 6ka in Midwestern USA, Global and

Planetary Change, 28, 285-291 (2001).

Cande, S.C., Kent, D.V., Revised calibration of the geomagnetic polarity time scale for

the late Cretaceous and Cenozoic, Journal of Geophysical Research, 100, 6093-

6095 (1995).

Camuffo, D., Physical weathering of stones, The Science of the Total Environment 167,

1-14 (1995).

Coe, R.S., and Liddicoat, J.C., Overprinting of natural magnetic remanence in lake

sediments by a subsequent high-intensity field, Nature, 367, 57-59 (1994).

Cohen, A.S., Paleolimnology. Oxford University Press, New York, 500pp. 2003.

Dasher, G.R and Balfour W., The caves and karst of the Buckeye Creek Basin. West

Virginia Speleological Survey, Barracksville, WV., 283p (1994). 88 Dasher, G.R., and Boyer, D.G., Dye tracings in the Spring Creek area, Greenbrier

County, West Virginia. IN: Younos, T., Burbey, T.J., Kastning, E.H., Poff, J.A.,

Karst-Water Environment Symposium. Virginia Water Resources Research

Center, Blacksburg, VA, 23-36 (1997).

Dasher, G.R., and Boyer, D.G., Recent Spring Creek area dye tracings. West Virginia

Speleological Society Monograph #2, Color Craft Printing, Charleston, WV, 29p

(2000).

Denniston, R.F., González, L.A., Asmerom, Y., Baker, R.G., Reagen, M.L., Bettis III,

E.A., Evidence for increased cool season moisture during the middle Holocene,

Geology 27(9), 815-818 (1999).

Dorale, J.A., Edwards, R.L., Alexander Jr., E.C., Shen, C., Richards, D.D., Cheng, H.,

Uranium-Series dating of speleothems: current techniques, limits, and

applications, in: Sasowsky, I.D., Mylroie, J. (Eds.), Studies of Cave Sediments,

Kluwer Academic, New York, pp177-197 (2004).

Drever, J.I., The Geochemistry of Natural Waters, 3rd Ed. Prentice-Hall, Inc., Englewood

Cliffs, NJ, 436pp, (1997).

Droxler, A.W., and Farrell, J.W., Marine Isotope Stage 11 (MIS 11): new insights for a

warm future, Global and Planetary Change 24(1), 1-5 (2000).

Dyurgerov, M.B., Meier, M.F., Twentieth century climate change: Evidence from small

glaciers. Proceedings of the National Academy of Sciences of the United States of

America 97(4), 1406-1411 (2000). 89 Eaton, L.S., Morgan, B.A., Kochel, R.C., Howard A.D., Quaternary deposits and

landscape evolution of the central Blue Ridge of Virginia, Geomorphology 56(1-

2), 139-154 (2003).

Einsele, G., Hinderer, M., Quantifying denudation and sediment-accumulation systems

(open and closed lakes): basic concepts and first results, Palaeogeography,

Palaeoclimatology, Palaeoecology, 140, 7-21 (1998).

Ely, L., Response of extreme floods in the southwest United States to climatic variations

in the late Holocene. Geomorphology, 19, 175-201 (1997).

Fairbanks, R.G., and Matthews, R.K., The Marine Oxygen Isotope Record in Pleistocene

Coral, Barbados, West Indies, Quaternary Research 10, 181-196 (1978).

Ford, D., Williams, P., Karst Geomorphology and Hydrology. Cambridge University

Press, Cambridge, Great Britain, 601p (1989).

Ford, D.C., Williams, P., Karst Hydrogeology and Geomorphology, Revised., Wiley,

New York, 576 pp. (2007).

Gascoyne, M., Palaeoclimate determination from cave calcite deposits, Quaternary

Science Reviews, 11, 609-632 (1992).

Gooch, J.L., Hetrick, S.W., The Relation of Genetic Structure to Environmental

Structure: Gammarus minus in a Karst Area. Evolution 33(1), 192-206 (1979).

Gordon, N.D., McMahon, T.A., Finlayson, B.L., Gippel, C.J., Nathan, R.J., Stream

Hydrology: An Introduction for Ecologists, 2nd Ed. John Wiley & Sons Ltd.

Hoboken, NJ, 429pp, (2004). 90 Granger, D.E., Fabel, D., Palmer, A.N., Pliocene-Pleistocene incision of the Green River,

Kentucky, determined from radioactive decay of cosmogenic 26Al and 10Be in

Mammoth Cave sediments. GSA Bulletin 113(7), 825-836 (2001).

Granger, D.E., Kirchner, J.W., Finkel, R.C., Quaternary downcutting rate of the New

River, Virginia, measured from differential decay of cosmogenic 26Al and 10Be in

cave-deposited alluvium. Geology 25, 107-110 (1997).

Hajna, N.Z., Mihevc, A., Pruner, P., Bosák, P., Palaeomagnetism and

magnetostratigraphy of karst sediments in Slovenia, ZRC Publishing, Ljubljana,

267pp, 2008.

Han, G., and Liu, C.-Q., Water geochemistry controlled by carbonate dissolution: a study

of the river waters draining karst-dominated terrain, Guizhou Province, China.

Chemical Geology 204, 1-21 (2004).

Hancock, G., Kirwan, M., Summit erosion rates deduced from 10Be: Implications for

relief in the central Appalachians. Geology 35(1), 89-92 (2007).

Hocutt, C.H., Denoncourt, R.F., Stauffer, Jr., J.R., Fishes of the Greenbrier River, West

Virginia, with Drainage History of the Central Appalachians. Journal of

Biogrography 5(1), 59-80 (1978).

Houghton J.T., Ding Y., Griggs D.J., Noguer M., van der Linden P.J., Dai X., Maskell

K., Johnson C.A. (Editors): Climate Change 2001: The Scientific Basis,

Contribution of Working Group 1 to the Third Assessment Report of the

Intergovernmental Panel on Climate Change, Cambridge University Press,

Cambridge, 892pp. (2001). 91 Houser, B.B., Erosional History of the New River, southern Appalachians, U.S.

Geological Survey Open-File Report 81-771, 225pp (1981).

Howard, A.D., Long Profile Development of Bedrock Channels: Interaction of

Weathering, Mass Wasting, Bed Erosion, and Sediment Transport, Rivers Over

Rock: Fluvial Processes in Bedrock Channels, Geophysical Monograph 107.

American Geophysical Union, 297-319 (1998).

Howard, A.D., Dietrich, W.E., Seidl, M.A., Modeling fluvial erosion on regional to

continental scales, Journal of Geophysical Research 99(B7), 971-986 (1994).

Imbrie, J., Boyle, E.A., Clemens, S.C., Duffy, A., Howard, W.R., Kukla, G., Kutzbach,

J., Martinson, D.G., McIntyre, A., Mix, A.C., Molfino, B., Morley, J.J., Peterson,

L.C., Pisias, N.G., Prell, W.L., Raymo, M.E., Shackleton, N.J., Toggweiler, J.R.,

On the structure and origin of major glaciation cycles 1. Linear responses to

Milankovitch forcing, Paleoceanography 7(6), 701-738 (1992).

Jones, W.K., Hydrogeology of Limestone Karst in Greenbrier County, West Virginia.

U.S. Geological Survey in association with the West Virginia Geological and

Economic Survey, Bulletin 36, 49p (1973).

Johnson, P. A., Gomez B., Cave levels and cave development in the Mitchell plain

following base-level lowering. Earth Surface Processes and Landforms, 19(6),

517-524 (1994).

Kines, J., Kleiber, H.-P., Matthiessen, J., Müller, C., Nowaczyk, N., Marine ice-rafted

debris records constrain maximum extent of Saalian and Weichselian ice-sheets

along the northern Eurasian margin, Global and Planetary Change 31, 45-64

(2001). 92 Kirchner, J.W., Dietrich, W.E., Iseya, F., Ikeda, H., The variability of critical shear stress,

friction angle, and grain protrusion in water-worked sediments, Sedimentology 37,

647-672 (1990).

Kump, L.R., Brantley, S.L., Arthur, M.A., Chemical weathering, atmospheric CO2, and

climate. Annu. Rev. Earth Planet Sci.28, 611-667 (2000).

Laird, K.R., Fritz, S.C., Grimm, E.C., Meuller, P.G., Century-scale paleoclimatic

reconstruction from Moon Lake, a closed lake basin in the Northern ,

Limnology and Oceanography, 41, 890-902 (1996).

Langmuir, D., Uranium solution-mineral equilibria at low temperatures with applications

to sedimentary ore deposits, Geochim. Cosmochim. Acta 42, 547-569 (1978).

Langmuir, D., and Herman, J.S., The mobility of thorium in natural waters at low

temperatures, Geochim. Cosmochim. Acta 44, 1753-1766 (1980).

Lea, D.W., Pak, D.K., Spero, H.J., Climate Impact of Late Quaternary Equatorial Pacific

Sea Surface Temperature Variations, Science 289, 1719-1724 (2000).

Leigh, D.S., Late Quaternary climates and river channels of the Atlantic Coastal Plain,

Southeastern USA. Geomorphology 101(1-2), 90-108 (2008).

Liu, W., Huang, Y., An, Z., Clemens, S.C., Li, L., Prell, W.L., Ning, Y., Summer

monsoon intensity controls C4/C3 plant abundance during the last 35 ka in the

Chinese Loess Plateau: Carbon isotope evidence from bulk organic matter and

individual leaf waxes, Palaeogeography, Palaeoclimatology, Palaeoecology 220,

243-254 (2005).

Lisiecki, L.E., Raymo, M.E., Plio-Pleistocene climate evolution: trends and transitions in

glacial cycle dynamics. Quaternary Science Reviews 26(1-2), 56-69 (2007). 93 Løvile, R., Sandnes, A., Palaeomagnetic excursions recorded in min-Weichselian cave

sediments from Skjonghelleren Valderøy, W. Norway, Physics of the Earth and

Planetary Interiors, 45, 337-348 (1987).

Ludwig, K.R., Simmons, K.R., Szabo, B.J., Winograd, I.J., Landwehr, J.M., Riggs, A.C.,

Hoffman, R.J., Mass-spectrometric 230Th-234U-238U dating of the Devils Hole

calcite vein, Science 258 (5080), 284-287 (1992).

Mackay, A., Battarbee, R., Birks, J., Oldfield., Global Change in the Holocene. Hodder

Arnold. Oxford University Press, 528 (2005).

Matmon, A., Bierman P.R., Larsen, J., Southworth, S., Pavich, M., Finkel, R., Caffee, M.,

Erosion of an Ancient Range, The Great Smoky , North

Carolina and Tennessee. American Journal of Science 303(9), 817-855 (2003).

McDermott, F., Palaeo-climate reconstruction from stable isotope variations in

speleothems: a review. Quaternary Science Reviews 23, 901-918 (2004).

Meyers, P.A., and Lallier-Vergès, E., Lacustrine sedimentary organic records of Late

Quaternary paleoclimates, Journal of Paleolimnology 12, 345-372 (1999).

Mills, H. H., Apparent increasing rates of stream incision in the eastern United States

during the late Cenozoic. Geology 28(10), 955-957 (2000).

Mills, H.H., Brakenridge, G.R., Jacobson, R.B., Newell, W.L., Pavich, M.J., Pomeroy,

J.S., and plateaus. In: Graff, W.L. (Ed.), Geomorphic

Systems of North America. Geological Society of America Centennial Special, 2,

5-50 (1987).

Monroe, J.S., and Wicander, R., The Changing Earth, 4th Ed. Thomson Brooks/Cole,

Courier Corporation/Kendallville, 754pp, (2006). 94 Montgomery, D.R., Abbe, T.B., Buffington, J.M., Peterson, N.P., Schmidt, K.M., Stock,

J.D. Distribution of bedrock and alluvial channels in forested mountain drainage

areas, Nature 381, 587-589 (1996).

Musgrave, R.J., and Webb, J.A., Palaeomagnetic analysis of sediments in the Buchan

Caves, Southeasters Australia, provides a pre- date for landscape

and climate evolutaion, in: Sasowsky, I.D., Mylroie, J. (Eds.), Studies of Cave

Sediments, Kluwer Academic, New York, pp47-69 (2004).

Nelson, K. J. P., Nelson, F. E. and Walegur, M. T. Periglacial Appalachia: palaeoclimatic

significance of blockfield elevation gradients, eastern USA. Permafrost and

Periglacial Processes 18(1), 61-73 (2007).

Ogden, A.E., Karst denudation rates of selected spring basins in West Virginia, NSS

Bulletin, 44, 6-10 (1982).

Palmer, A.N., Cave Geology, Watson, R.A. (Eds.), Allen Press, Kansas, 454 pp. 2007.

Palmer, A.N., Cave levels and their interpretation. National Speleological Society

Bulletin 49(2), 50-66 (1987).

Panno, S.V., Curry, B.B., Wang, H., Hackley, K.C., Liu, C-L., Lundstrom, C., Zhou, J.,

Climate change in southern Illinois, USA, based on the age and δ13C of organic

matter in cave sediments, Quaternary Research 61, 301-313 (2004).

Peterson, B.J., and Fry, B., Stable Isotopes in Ecosystem Studies, Annual Review of

Ecology and Systematics 18, 293-320 (1987).Ritter, D.F., Kochel, R.C., Miller,

J.R., Process Geomorphology, 4th Ed. Waveland Press, Inc. Long Grove, IL,

560pp, (2002). 95 Petit, J.R., Jouzel, J., Raynaud, D., Barkov, N.I., Barnola, J.-M., Basile, I., Bender, M.,

Chappellaz, J., Davis, M., Delaygue, G., Delmotte, M., Kotlyakov, V.M.,

Legrand, M., Lipenkov, V.Y., Lorius, C., Pépin, L., Ritz, C., Saltzman, E.,

Stievenard, M., Climate and atmospheric history of the past 420,000 years from

the Vostok ice core, Antarctica, Nature, 399, 429-436 (1999).

Potter, E.-K., and Lambeck, K., Reconciliation of sea-level observations in the Western

North Atlantic during the last glacial cycle, Earth and Planetary Science Letters

217, 171-181 (2003).

Price, P.H., Heck, E.T., West Virginia Geological Survey: Greenbrier County. Wheeling

News Lithograph Co., Wheeling WV., 846 (1939).

Reusser, L., Bierman, P., Pavich, M., Zen, E., Larsen, J., Finkel, R., Rapid Late

Pleistocene Incision of Atlantic Passive-Margin River Gorges. Science 305, 499-

502 (2004).

Robinson, L.F., Henderson, G.M., Slowey, N.C., U-Th dating of marine isotope stage 7

in Bahamas slope sediments, Earth and Planetary Science Letters 196(3-4), 175-

187 (2002).

Sasowsky, I.D., Paleomagnetic record in cave sediments, in: Culver D.C. and White,

W.B. (Eds.), Encyclopedia of caves: San Diego, Elsevier / Academic Press,

pp427-431 (2005).

Sasowsky, I.D., Clotts, R.A., Crowell, B., Walko, S.M., LaRock, E.J., Harbert, W.,

Paleomagnetic analysis of a long-term sediment trap, Kooken Cave, Huntingdon

County, Pannsylvania, USA, in: Sasowsky, I.D., Mylroie, J. (Eds.), Studies of

Cave Sediments, Kluwer Academic, New York, pp71-81 (2004). 96 Sasowsky, I.D., White, W.B., Schmidt, V.A., Determination of stream-incision rate in the

Appalachian plateaus by using caves-sediment magnetostratigraphy, Geology,

23(5), 415-418 (1995).

Schmidt, V. A., Magnetostratigraphy of Sediments in Mammoth Cave, Kentucky.

Science, 217(4562), 827-829 (1982).Shackleton, N.J., Sánchez-Goñi, M.F.,

Pailler, D., Lancelot, Y., Marine Isotope Substage 5e and the Eemian Interglacial,

Global and Planetary Change 36, 151-155 (2003).

Schmidt, V.A., Jennings, J.N., Haosheng, B., Dating of cave sediments at Wee Jasper,

New South Wales, by magnetostratigraphy, Australian Journal of Earth Sciences,

31, 361-370 (1984).

Schneider, K., and Culver, D.C., Estimating subterranean species richness using intensive

sampling and rarefaction curves in a high density cave region in West Virginia,

Journal of Cave and Karst Studies 66(2), 39-45 (2004).

Springer, G.S., A pipe-based, first approach to modeling closed conduit flow in caves.

Journal of Hydrology 289(1-4), 178-189 (2004).

Springer, G.S., Caves and Their Potential Use in Paleoflood Studies, In: Ancient Floods,

Modern Hazards: Principles and Applications of Paleoflood Hydrology, Water

Science and Application, 5, House, P.K., Webb, R.H., Baker, V.R., Levish, D.R.,

eds. Washington D.C., American Geophysical Union, 329-343 (2002b).

Springer, G.S., Clastic Sediments in Caves. In: Encyclopedia of Caves (Culver, D., and

White, W., Editors), Elsevier, ISBN: 0-12-198-651-9, 102-108 (2005).

Springer, G.S., Stream Processes in Cave Surface Streams flowing over Limestone,

Southern Friends of the Pleistocene Fall 2002 Fieldtrip Guidebook, West 97 Virginia Speleological Society Monograph #3, Ohio University, Athens, OH,

109pp (2002a).

Springer, G.S., Dowdy, H.S., Eaton, L.S., Sediment budgets for two mountainous basins

affected by a catastrophic storm: Blue Ridge Mountains, Virginia,

Geomorphology 37, 135-148 (2001).

Springer, G.S., Kite, J.S., Schmidt, V.A., Cave sedimentation, genesis, and erosional

history in the Cheat River Canyon, West Virginia. GSA Bulletin 109(5), 524-532

(1997).

Springer, G.S., Rowe, H., Hardt, B., Cocina, F.G., Edwards, R.L., and Cheng H.,

Climate Driven Changes In River Channel Morphology And Base Level During

The Holocene And Late Pleistocene Of Southeastern West Virginia. Journal of

Cave and Karst Studies 71(2), (2009).

Springer, G.S., Rowe, H.D., Hardt, B., Edwards, R.L., Chend, H., Solar forcing of

Holocene droughts in a stalagmite record from West Virginia in east-central

North America, Geophysical Research Letters, 35, L17703 (2008)

Springer, G.S., White, D.M., Rowe, H.D., Hardt, B., Mihimdukulasooriya, L.N., Cheng,

Edwards, R.L., Multiproxy evidence from caves of Native Americans altering the

overlying landscape during the late Holocene of east-central North America, The

Holocene 20(2), 275-283 (2010).

Springer, G.S., and Whol, E.E., Empirical and Theoretical Investigations of Sculpted

Forms in Buckeye Creek Cave, West Virginia, The Journal of Geology 110, 469-

481 (2002). 98 Springer, G.S., Wohl, E.E., Foster, J.A., Boyer D.G., Testing for reach-scale adjustments

of hydraulic variables to soluble and insoluble strata: Buckeye Creek and

Greenbrier River, West Virginia. Geomorphology 56(1-2), 201-217 (2003).

Stirling, C.H., Esat, T.M., Lambeck, K., McCulloch, M.T., Blake, S.G., Lee, D.-C.,

Halliday, A.N., of the Marine Isotope Stage 9 Interglacial,

Science 291, 290-293 (2001).

Stock, G. M., Anderson, R.S., Finkel, R.C., Rates of erosion and topographic evolution of

the Sierra Nevada, California, inferred from cosmogenic 26Al and 10Be

concentrations. Earth Surface Processes and Landforms 30(8), 985-1006 (2005a).

Stock, G.M., Granger, D.E., Sasowsky, I.D., Anderson, R.S., Finkel, R.C., Comparison

of U-Th, paleomagnetism, and cosmogenic burial methods for dating caves:

Implications for landscape evolution studies. Earth and Planetary Science Letters

236, 388-403 (2005b).

Taylor, S.B., and Kite, S.J., Comparative geomorphic analysis of surficial deposits at

three central Appalachian watersheds: Implications for controls on sediment-

transport efficiency, Geomorphology 78, 22-43 (2006).

Tebbens, L.A., Veldkamp, A., Westerhoff, W. and Kroonenberg, S.B., Fluvial incision

and channel downcutting as a response to Late-glacial and Early Holocene

climate change: the lower reach of the River Meuse (Maas), The .

Journal of Quaternary Science 14, 59–75 (1999).

Tucker, G.E., and, Bras, R.L., Hillslope processes, drainage density, and landscape

morphology. Water Resources Research 34(10), 2751-2764 (1998). 99 Tucker, G.E., and Slingerland, R., responses to climate change. Water

Resources Research 33(8), 2031-2047 (1997).

Vandenberghe, J., Timescales, Climate, and River Development, Quaternary Science

Reviews, 14, 631-638 (1995).

Verosub, K.L., Depositional and postdepositional processes in the magnetization of

sediments. Reviews of Geophysics and Space Physics, 15(2), 129-143 (1977).

Ward, D.J., Spotila, J.A., Hancock, G.S., Galbraith, J.M., New constraints on the late

Cenozoic incision history of the New River, Virginia. Geomorphology 72, 54-72

(2005).

Wegmann, K.W., Pazzaglia, F.J., Holocene terraces, climate change, and active

tectonics: The Clearwater River basin, Olympic Peninsula, Washington State.

GSA Bulletin 114(6), 731-744 (2002)

Werner, E., Hydrochemical Characteristics of the Greenbrier Limestone Karst of East-

Central West Virginia, in Kastning E.H. and Kastning K.M. (Eds.), Appalachian

Karst, National Speleological Society, Huntsville, AL, pp187-196 (1991).

Wendland, W.M., Climate Changes: Impacts on Geomorphic Processes. Engineering

Geology, Elsevier Science, 45, 347-358 (1996).

White, D.M., Reconstruction and analysis of Native American land use during the Late

Holocene (M.S. Thesis): Athens, Ohio, Ohio University, 164 pp. (2007).

White, W.B., Paleoclimate records from speleothems in limestone caves, in: Sasowsky,

I.D., Mylroie, J. (Eds.), Studies of Cave Sediments, Kluwer Academic, New

York, pp135-175 (2004). 100 White, W.B., Cave sediments and paleoclimate. Journal of Cave and Karst Studies,

69(1), 76-93 (2007).

Wilby, R.L., Dalgleish, H.Y., Foster, D.L., The Impact of Weather Patterns on Historic

and Contemporaneous Catchment Sediment Yields. Earth Surface Processes and

Landforms, 22, 353-363 (1997).

Worthington, S.R.H., The paleodrainage of an Appalachian fluviokarst: Friars Hole, West

Virginia (M.S. Thesis): Hamilton, Ontario, McMaster University, 218p (1984).

Xu, Z., and Liu, C.-Q., Chemical weathering in the upper reaches of Xijiand River

draining the Yunnan – Guizhou Plateau, Southwest China. Chemical Geology

239, 83-95 (2007).

Yoshimura, K., Nako, S., Noto, M., Inokura, Y., Urata, K., Chen, M., Lin, P.-W.,

Geochemical and stable isotope studies on natural water in the Taroko Gorge

karst area, Taiwan – chemical weathering of carbonate rocks by deep source CO2

and sulfuric acid. Chemical Geology 177, 415-430 (2001).

Zhaoyan, G., Qiang, L., Bing, X., Jiamao, H., Shiling, Y., Zhongli, D., Tungsheng, L.,

Climate as the dominant control on C3 and C4 plant abundance in the Loess

Plateau: Organic carbon isotope evidence from the last glacial-interglacial loess-

soil sequences, Chinese Science Bulletin 48 (12), 1271-1276 (2003).

101 8.0 Appendix

8.1 Appendix A: Survey Data

Setup July #1 30 Dist Inc Relative to Absolute above rate Station Notes (m) (degrees) instrument z (m) stream (m/Ma) Stalagmite in phreatic loop 1 1.46 13.4 0.338 100.944 18.293 top Top of 2 1.13 -33.2 -0.619 99.987 17.336 cobbles A3 from original 3 4.90 -7.1 -0.606 100.000 17.349 survey Cobbles 4 5.62 10.1 0.986 101.591 18.940 above A3 Solutional ceiling above the face rock where we change 5 10.71 9.7 1.805 102.410 19.759 clothes Cobble fill above 6 10.30 3.3 0.593 101.199 18.547 phreatic loop Location of BCC-27 stalagmite, 7 3.14 -32.6 -1.692 98.914 16.263 539000ka 30.172

8 15.08 -4.8 -1.262 99.344 16.693 To 2-1

Setup #2

Station

1 8.09 -4.4 -0.621 99.344 16.693 To 1-8 To the tall "trophy" 2 6.36 -11.6 -1.279 98.686 16.034 stalagmite 102 Flat patch on ceiling above breakdown 3 12.35 12.3 2.631 102.595 19.944 blocks

4 25.34 1.2 0.531 100.495 17.844 To 3-1

Setup #3

Station

1 1.84 13.3 0.423 100.495 17.844 To 2-4

2 18.55 -0.2 -0.065 100.007 17.356 To 4-1

Setup #4

Station

1 7.13 -8.7 -1.078 100.007 17.356 To 3-2 Ceiling 2 7.15 11.9 1.474 102.560 19.909 above hole Stalagmite semi-circle, 3 1.89 -19.5 -0.631 100.455 17.803 far left Silt above 4 7.22 0.7 0.088 101.174 18.523 cobble Cobble top 5 7.01 -0.4 -0.049 101.037 18.385 below silt

6 18.72 -3.7 -1.208 99.878 17.226 To 5-1

Setup #5

Station 103

1 3.52 -5.1 -0.313 99.878 17.226 To 4-6 Rock in front of trench to 2 7.17 4.3 0.538 100.728 18.077 6-1

3 6.29 18.7 2.017 102.207 19.556 Top of silt Ceiling towards 4 16.56 20.1 5.691 105.881 23.230 trench Ceiling back towards setup 5 9.80 15.5 2.619 102.809 20.158 4

Setup #6

Station To 5-2 (real shot should be 2cm 1 19.32 -2.5 -0.843 100.728 18.077 higher) Top of silt near snail location, actual top is 2 8.64 10.3 1.545 103.116 20.464 ~8cm higher Top of gravel to right of 3 8.21 4.7 0.673 102.243 19.592 turn Top of gravel, actual top is 2cm 4 10.65 4.5 0.836 102.406 19.755 higher Bedrock below 5 10.14 -0.3 -0.053 101.518 18.866 cobbles Ceiling where passages 6 11.42 7.4 1.471 103.042 20.390 connect 104 Edge of water down passage (incorrect 7 77.07 -25.7 -33.422 68.149 -14.503 measurement from before Dr.S went down to stream) 4cm higher than the edge of the water down passage to stream (high 8 43.47 -25.8 -18.919 82.651 0.000 flow)

Setup Oct #7 17 To 5-2/6-1 (rock in front 1 0.94 0.2 0.003 100.728 18.077 of trench) Up passage to right of trench, towards 2 12.74 18.8 4.106 104.830 22.179 Stope's Top, breakdown block that fell on top of stalagmite

Setup #8

Station To breakdown 1 1.96 -6.9 -0.235 104.830 22.179 block 7-2 Up stope to mud cairn, up 2 17.69 24.5 7.336 112.402 29.751 passage to 105 left is a block sticking out

Setup #9

Station To mud cairn 1 2.59 -8.4 -0.378 112.402 29.751 8-2 Reversed sediments down the hole, 2 2.10 -38.6 -1.310 111.470 28.819 773000ka 37.282 Gravel and cobble unit above hole, above reversed sediments, 3 1.37 -18.9 -0.444 112.336 29.685 FUS Sand and gravel unit above 9-4, fining upward 4 14.40 15.9 3.945 116.725 34.074 sequence Gravels and sands at base of third fining upward 5 1.50 30.2 0.755 113.535 30.883 sequence Ceiling up stope from fining upward 6 5.15 32.5 2.767 115.547 32.896 sequences Solutional ceiling above mud cairn - 7 2.98 90.0 2.980 115.760 33.109 straight up

106 Setup #10

Station Stalagmite in phreatic loop 1 1.36 -7.0 -0.166 100.944 18.293 from 1-1 Solutional ceiling of phreatic loop looking down towards 2 11.15 -10.6 -2.051 99.059 16.407 active stream Top of chert layer along side of 3 9.98 -14.0 -2.414 98.695 16.044 phreatic loop Rock (breakdown block) in side of loop - possibly A2 4 13.31 -18.4 -4.201 96.908 14.257 from original survey, the survey mark is unclear could also be cc1.

Setup #11

Station Back to 10-4, 1 1.40 17.9 0.430 96.908 14.257 rock in loop Breakdown ceiling edge 2 11.68 -3.0 -0.611 95.867 13.216 down loop Solutional ceiling - shooting lowest part of 3 21.69 -9.3 -3.505 92.973 10.322 orifices Solutional 4 26.45 -10.1 -4.638 91.840 9.188 ceiling in 107 loop

Solutional ceiling in 5 30.36 -10.2 -5.376 91.102 8.451 loop Solutional ceiling in 6 36.94 -11.0 -7.048 89.430 6.778 loop Solutional ceiling of downstream stream 7 41.87 -14.0 -10.129 86.349 3.698 passage Waters edge, actual reading is ~4cm below 8 49.24 -16.3 -13.820 82.658 0.007 this Down tunnel left of 10- 9 15.93 -1.9 -0.528 95.950 13.299 4/11-1

Setup #12

Station

1 2.13 12.0 0.443 95.950 13.299 To 11-9 Top of stalagmite in passage left 2 11.80 8.3 1.703 97.211 14.559 of 11-9/12-1

Setup #13

Station Back to stalagmite 1 7.46 1.8 0.234 97.211 14.559 top 12-2 Pendant on solutional 2 6.20 16.9 1.802 98.779 16.127 ceiling 108 Top of cobble unit, actual top is ~4cm higher, 3 18.83 2.6 0.854 97.830 15.179 this unit may have gone up to the ceiling and slumped down Top of 4 15.94 7.3 2.025 99.002 16.350 stalagmite Solutional ceiling - continuation 5 21.30 17.0 6.228 103.204 20.552 of BRT

Setup #14

Station Top of stalagmite 1 0.95 -24.2 -0.389 99.002 16.350 from 13-4 Top of solutional ceiling near 2 11.76 -4.0 -0.820 98.571 15.919 column Former location of stalagmite taken for 3 16.04 -12.2 -3.390 96.001 13.350 dating, floor of ceiling shot in 14-3, near column Wall, continuation 4 20.43 5.1 1.816 101.207 18.556 of BRT

Solutional 5 18.64 12.1 3.907 103.298 20.647 ceiling

Setup #15 109

Station

1 3.80 -7.4 -0.489 101.207 18.556 Back to 14-4 Base of wall notch, may be continuous with cobble 2 2.28 2.8 0.111 101.808 19.157 unit "Top" of notch, part of solutional ceiling shot 3 2.50 13.3 0.575 102.272 19.620 in 14-5 Station ?2 of original 4 5.52 2.6 0.250 101.947 19.296 survey Rock Dr. Springer put 5 12.47 0.1 0.022 101.718 19.067 in the ground

Setup #16

Station Back to rock 1 4.55 -1.3 -0.103 101.718 19.067 15-5 To short 2 14.64 -3.0 -0.766 101.055 18.404 crawl-way

3 5.82 16.5 1.653 103.475 20.823 Ceiling

Setup #17

Station Back to point in crawl-way 1 2.66 -11.4 -0.526 101.055 18.404 16-2 110 Piece of popcorn on 2 13.80 -1.9 -0.458 101.124 18.472 wall rock

Setup #18

Station Back to 1 3.35 26.7 1.505 101.124 18.472 popcorn 17-2 Cobbles (have slumped), scallops indicate flow is in the expected direction, from setup 17 over cobbles 2 2.51 9.3 0.406 100.024 17.373 to setup 18 ?6 from original 3 2.23 -2.3 -0.089 99.529 16.878 survey

Breakdown 4 18.15 11.8 3.712 103.330 20.679 block

Setup #19

Station Back to breakdown 1 6.56 -14.8 -1.676 103.330 20.679 block 18-4 Tip of rock in 2 5.69 6.0 0.595 105.600 22.949 cobbles ?9, top of 3 8.30 2.6 0.377 105.382 22.731 stalagmite

4 18.17 8.1 2.560 107.566 24.915 Wall point 111

Setup #20

Station Wall point 1 3.39 4.8 0.284 107.566 24.915 19-4 BCC 21 & 22 location, BCC-22 : 2 5.23 10.0 0.908 108.190 25.539 539925ka 47.301 BCC 25 location : 3 6.63 -4.4 -0.509 106.774 24.122 488777ka 49.352 Solutional ceiling for 4 5.57 24.6 2.319 109.601 26.950 prism canyon Approximate contact between limestone and prismatic 5 7.75 10.9 1.465 108.748 26.096 jointing unit Solutional ceiling upper 6 15.76 8.2 2.248 109.530 26.879 tube BRT 7 13.68 -11.3 -2.681 104.602 21.950 connection

Setup #21

Station To 19-2, tip of rock in 1 2.29 8.5 0.338 105.600 22.949 cobbles Top of 2 19.77 -0.4 -0.138 105.124 22.473 stalagmite

Setup #22 112

Station Top of stalagmite, actual top is 1 14.71 2.4 0.616 105.124 22.473 1.5cm higher BCC 13 & 14 location, BCC-13 : 2 9.01 -7.9 -1.238 103.270 20.618 293936ka 70.146 Ceiling, probably 3 14.37 8.1 2.025 106.533 23.881 breakdown Top of 4 31.21 -1.7 -0.926 103.582 20.931 stalagmite Top of silt behind 5 25.22 2.9 1.276 105.784 23.133 speleothems

Setup #23

Station Back to stalagmite 1 8.39 13.3 1.930 103.582 20.931 22-4 Pointy breakdown 2 6.58 15.8 1.792 103.444 20.792 piece Top of stream 3 8.80 14.4 2.188 103.840 21.189 sediments

Setup #24

Station

1 1.72 9.5 0.284 103.444 20.792 To 23-2 ?101? original survey point, 2 15.03 -13.8 -3.585 99.575 16.923 actual point 113 is 2cm to right Lumpy boulder on top of large breakdown 3 22.24 -6.0 -2.325 100.835 18.184 pile

Setup #25

Station To 24-3 lumpy boulder on top of large breakdown 1 8.14 15.3 2.148 100.835 18.184 pile Ceiling for BCC2 2 14.46 8.4 2.112 100.799 18.148 passage Down passage boulder 3 16.53 -24.4 -6.829 91.858 9.207 breakdown

Setup #26

Station To 25-3 boulder 1 6.19 -2.8 -0.302 91.858 9.207 breakdown Bottom of pendants, lots of anastamose 2 9.61 13.4 2.227 94.388 11.737 tubes Ceiling tube (half-tube), solutional ceiling for 3 21.84 19.1 7.146 99.307 16.656 the big trunk 114

4 30.37 8.8 4.646 96.807 14.156 Ceiling ledge

Setup #27

Station Ceiling ledge 1 6.19 29.0 3.001 96.807 14.156 26-4 Continuation of ceiling tube (half- 2 9.11 32.9 4.948 98.754 16.103 tube)

3 26.62 -0.8 -0.372 93.434 10.783 Wall point

Setup #28

Station Top of 1 6.42 34.3 3.618 96.973 14.321 cobbles Bottom of 2 5.82 28.1 2.741 96.096 13.445 cobbles

3 2.07 2.2 0.079 93.434 10.783 Wall 27-3 D7 original 4 10.67 2.0 0.372 93.727 11.076 survey

Setup #29

Station D7 original 1 2.08 11.3 0.408 93.727 11.076 & 28-4 Contact between flat 2 3.31 31.6 1.734 95.054 12.403 bed and bed 115 that is squiggly from paragenesis Bottom of cobbles on opposite wall 3 5.79 37.3 3.509 96.828 14.177 of 29-2

4 10.10 37.7 6.176 99.496 16.845 Ceiling tube

Setup Oct #30 18 D7 original 1 1.81 13.0 0.407 93.727 11.076 & 28-4/29-1 Breakdown 2 26.61 -2.2 -1.021 92.299 9.647 block

Setup #31

Station Breakdown 1 3.73 -11.6 -0.750 92.299 9.647 block 30-2 Chert nodules in ceiling, back towards 2 10.37 28.4 4.932 97.981 15.330 station 30 Bottom of cobbles on 3 13.54 15.3 3.573 96.621 13.970 left wall Top of 4 14.53 18.3 4.562 97.611 14.960 cobbles F3 original 5 28.47 1.7 0.845 93.893 11.242 survey

Setup #32 116

Station F3 original 1 2.07 -9.9 -0.356 93.893 11.242 survey

2 23.93 9.1 3.785 98.034 15.383 Wall point

Setup #33

Station Wall point, passage right, 1 1.57 25.3 0.671 98.034 15.383 old point ? Solutional 2 8.24 12.7 1.812 99.174 16.523 ceiling Top of stalagmite, tall one in front of 3 9.37 -2.9 -0.474 96.889 14.238 column

Setup #34

Station Top of stalagmite 1 10.77 -11.7 -2.184 96.889 14.238 33-3 Top of stalagmite by breakdown block 14- 2 2.90 11.3 0.568 99.641 16.990 1/13-4

117 8.2 Appendix B: Sieving Results

Sample: 3/1/09 HF_01

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 900 42.55 57.45

0.4380 11.125 -3.476 160 7.56 49.88

0.3120 7.925 -2.986 120 5.67 44.21

0.2230 5.664 -2.502 140 6.62 37.59

0.1570 3.988 -1.996 120 5.67 31.92

0.1100 2.794 -1.482 125.33 5.93 25.99

0.0787 1.999 -0.999 94.85 4.48 21.51

0.0555 1.410 -0.495 93.16 4.40 17.10

0.0394 1.001 -0.001 65.96 3.12 13.98

0.0278 0.706 0.502 46.85 2.22 11.77

0.0197 0.500 0.999 47.73 2.26 9.51

0.0139 0.353 1.502 40.34 1.91 7.60

0.0098 0.249 2.006 34.85 1.65 5.96

0.0070 0.178 2.492 55.38 2.62 3.34

0.0049 0.124 3.006 27.47 1.30 2.04

0.0025 0.006 7.310 24.45 1.16 0.88

18.7 0.88 0.00

Total = 2115.1

d16= 1.3

d50= 11.125

d84= 118

Sample: 3/1/09 HF_02

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 108.66 20.65 79.35

0.0394 1.001 -0.001 82.71 15.72 63.62

0.0278 0.706 0.502 55.34 10.52 53.10

0.0197 0.500 0.999 40.17 7.64 45.47

0.0139 0.353 1.502 27.36 5.20 40.27

0.0098 0.249 2.006 29.66 5.64 34.63

0.0070 0.178 2.492 15.8 3.00 31.63

0.0049 0.124 3.006 23.65 4.50 27.13

0.0025 0.006 7.310 67.18 12.77 14.36

75.56 14.36 0.00

Total = 526.09

d16=

d50= 0.6

d84= 1.7

119 Sample: 3/1/09 HF_03

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 102.59 21.76 78.24

0.0394 1.001 -0.001 40.63 8.62 69.62

0.0278 0.706 0.502 40.91 8.68 60.94

0.0197 0.500 0.999 36.26 7.69 53.25

0.0139 0.353 1.502 27.16 5.76 47.49

0.0098 0.249 2.006 46.12 9.78 37.71

0.0070 0.178 2.492 77.1 16.35 21.35

0.0049 0.124 3.006 60.87 12.91 8.44

0.0025 0.006 7.310 26.55 5.63 2.81

13.25 2.81 0.00

Total = 471.44

d16= 0.16

d50= 0.4

d84= 1.6 120

Sample: HF_09_001 Stope's Top, 189-186cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 6.72 4.78 95.22

0.3120 7.925 -2.986 9.89 7.04 88.18

0.2230 5.664 -2.502 13.39 9.53 78.65

0.1570 3.988 -1.996 9.06 6.45 72.20

0.1100 2.794 -1.482 5.99 4.26 67.93

0.0787 1.999 -0.999 3.8 2.70 65.23

0.0555 1.410 -0.495 3.3 2.35 62.88

0.0394 1.001 -0.001 8.65 6.16 56.72

0.0278 0.706 0.502 6.1 4.34 52.38

0.0197 0.500 0.999 5.87 4.18 48.20

0.0139 0.353 1.502 4.26 3.03 45.17

0.0098 0.249 2.006 10.24 7.29 37.88

0.0070 0.178 2.492 13.29 9.46 28.42

0.0049 0.124 3.006 15.51 11.04 17.38

0.0025 0.006 7.310 15.27 10.87 6.51

9.15 6.51 0.00

Total = 140.49

d16= 0.13

d50= 0.565

d84= 7 121

Sample: HF_09_002, 174-167cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 13.97 6.39 93.61

0.3120 7.925 -2.986 22.56 10.31 83.30

0.2230 5.664 -2.502 25.81 11.80 71.50

0.1570 3.988 -1.996 22.88 10.46 61.04

0.1100 2.794 -1.482 19.4 8.87 52.18

0.0787 1.999 -0.999 12.21 5.58 46.59

0.0555 1.410 -0.495 8.73 3.99 42.60

0.0394 1.001 -0.001 4.71 2.15 40.45

0.0278 0.706 0.502 10.13 4.63 35.82

0.0197 0.500 0.999 7.56 3.46 32.36

0.0139 0.353 1.502 6.89 3.15 29.21

0.0098 0.249 2.006 9.42 4.31 24.91

0.0070 0.178 2.492 16.61 7.59 17.32

0.0049 0.124 3.006 14.78 6.76 10.56

0.0025 0.006 7.310 11.91 5.44 5.12

11.19 5.12 0.00

Total = 218.76

d16= 0.17

d50= 2.5

d84= 8

122

Sample: HF_09_003,152cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0.67 0.39 99.61

0.2230 5.664 -2.502 4.33 2.54 97.07

0.1570 3.988 -1.996 3.38 1.98 95.09

0.1100 2.794 -1.482 5.63 3.30 91.80

0.0787 1.999 -0.999 3.89 2.28 89.52

0.0555 1.410 -0.495 3.56 2.08 87.43

0.0394 1.001 -0.001 4 2.34 85.09

0.0278 0.706 0.502 9.4 5.51 79.58

0.0197 0.500 0.999 12.12 7.10 72.49

0.0139 0.353 1.502 14.52 8.50 63.98

0.0098 0.249 2.006 26.92 15.77 48.22

0.0070 0.178 2.492 41.93 24.56 23.66

0.0049 0.124 3.006 27.41 16.05 7.61

0.0025 0.006 7.310 9.24 5.41 2.20

3.75 2.20 0.00

Total = 170.75

d16= 0.16

d50= 0.255

d84= 0.97 123

Sample: HF_09_004,281cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 3.92 2.26 97.74

0.3120 7.925 -2.986 6.27 3.62 94.12

0.2230 5.664 -2.502 7.65 4.42 89.70

0.1570 3.988 -1.996 6.53 3.77 85.93

0.1100 2.794 -1.482 10.24 5.91 80.01

0.0787 1.999 -0.999 10.87 6.28 73.74

0.0555 1.410 -0.495 11.27 6.51 67.23

0.0394 1.001 -0.001 21.93 12.66 54.56

0.0278 0.706 0.502 20.71 11.96 42.61

0.0197 0.500 0.999 14.63 8.45 34.16

0.0139 0.353 1.502 11.47 6.62 27.53

0.0098 0.249 2.006 10.47 6.05 21.49

0.0070 0.178 2.492 11.72 6.77 14.72

0.0049 0.124 3.006 10.61 6.13 8.59

0.0025 0.006 7.310 8.2 4.74 3.86

6.68 3.86 0.00

Total = 173.17

d16= 0.19

d50= 0.88

d84= 3.5

124 Sample: HF_09_005,267cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0.21 0.13 99.87

0.1570 3.988 -1.996 1.32 0.81 99.06

0.1100 2.794 -1.482 2.36 1.45 97.61

0.0787 1.999 -0.999 4.96 3.04 94.57

0.0555 1.410 -0.495 17.9 10.98 83.59

0.0394 1.001 -0.001 30.7 18.83 64.77

0.0278 0.706 0.502 36.19 22.19 42.57

0.0197 0.500 0.999 28.19 17.29 25.29

0.0139 0.353 1.502 16.37 10.04 15.25

0.0098 0.249 2.006 9.05 5.55 9.70

0.0070 0.178 2.492 6.39 3.92 5.78

0.0049 0.124 3.006 3.74 2.29 3.48

0.0025 0.006 7.310 2.7 1.66 1.83

2.98 1.83 0.00

Total = 163.06

d16= 0.353

d50= 0.8

d84= 1.41

125 Sample: HF_09_006,243cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0.58 0.50 99.50

0.1570 3.988 -1.996 1.59 1.38 98.12

0.1100 2.794 -1.482 5.26 4.56 93.55

0.0787 1.999 -0.999 8.51 7.38 86.17

0.0555 1.410 -0.495 20.35 17.66 68.51

0.0394 1.001 -0.001 25.5 22.13 46.38

0.0278 0.706 0.502 22.13 19.20 27.18

0.0197 0.500 0.999 11.33 9.83 17.35

0.0139 0.353 1.502 6.11 5.30 12.04

0.0098 0.249 2.006 4.69 4.07 7.97

0.0070 0.178 2.492 4.77 4.14 3.84

0.0049 0.124 3.006 2.46 2.13 1.70

0.0025 0.006 7.310 1.14 0.99 0.71

0.82 0.71 0.00

Total = 115.24

d16= 0.46

d50= 1.1

d84= 1.9

126 Sample: HF_09_007,225cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0.5 0.37 99.63

0.1570 3.988 -1.996 1.08 0.79 98.84

0.1100 2.794 -1.482 2.19 1.60 97.24

0.0787 1.999 -0.999 3.81 2.79 94.46

0.0555 1.410 -0.495 9.91 7.25 87.21

0.0394 1.001 -0.001 17.37 12.70 74.51

0.0278 0.706 0.502 23.63 17.28 57.23

0.0197 0.500 0.999 22.56 16.50 40.73

0.0139 0.353 1.502 18.92 13.84 26.90

0.0098 0.249 2.006 16.96 12.40 14.49

0.0070 0.178 2.492 7.72 5.65 8.85

0.0049 0.124 3.006 8.14 5.95 2.90

0.0025 0.006 7.310 2.55 1.86 1.03

1.41 1.03 0.00

Total = 136.75

d16= 0.275

d50= 0.6

d84= 1.3

127

Sample: HF_09_008

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 1.54 0.95 99.05

0.3120 7.925 -2.986 2.43 1.49 97.56

0.2230 5.664 -2.502 1.73 1.06 96.50

0.1570 3.988 -1.996 1.31 0.81 95.69

0.1100 2.794 -1.482 4.24 2.61 93.08

0.0787 1.999 -0.999 8.44 5.19 87.90

0.0555 1.410 -0.495 18.87 11.60 76.30

0.0394 1.001 -0.001 20.09 12.35 63.95

0.0278 0.706 0.502 21.5 13.22 50.73

0.0197 0.500 0.999 17.7 10.88 39.85

0.0139 0.353 1.502 14.37 8.83 31.02

0.0098 0.249 2.006 14.15 8.70 22.32

0.0070 0.178 2.492 15.12 9.29 13.03

0.0049 0.124 3.006 8.88 5.46 7.57

0.0025 0.006 7.310 5.97 3.67 3.90

6.34 3.90 0.00

Total = 162.68

d16= 0.2

d50= 0.69

d84= 1.8

128 Sample: HF_09_009,305cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 11.16 29.11 70.89

0.4380 11.125 -3.476 0 0.00 70.89

0.3120 7.925 -2.986 0 0.00 70.89

0.2230 5.664 -2.502 2.3 6.00 64.89

0.1570 3.988 -1.996 0.94 2.45 62.44

0.1100 2.794 -1.482 2.27 5.92 56.52

0.0787 1.999 -0.999 2.04 5.32 51.20

0.0555 1.410 -0.495 3.98 10.38 40.82

0.0394 1.001 -0.001 3.24 8.45 32.37

0.0278 0.706 0.502 2.55 6.65 25.72

0.0197 0.500 0.999 1.86 4.85 20.87

0.0139 0.353 1.502 1.54 4.02 16.85

0.0098 0.249 2.006 1.31 3.42 13.43

0.0070 0.178 2.492 1.36 3.55 9.89

0.0049 0.124 3.006 1.14 2.97 6.91

0.0025 0.006 7.310 1.16 3.03 3.89

1.49 3.89 0.00

Total = 38.34

d16= 0.31

d50= 1.9

d84=

129 Sample: HF_09_010,136cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0.03 0.02 99.98

0.0197 0.500 0.999 0.07 0.04 99.94

0.0139 0.353 1.502 0.75 0.48 99.46

0.0098 0.249 2.006 3.83 2.45 97.00

0.0070 0.178 2.492 30.76 19.70 77.30

0.0049 0.124 3.006 45.25 28.98 48.33

0.0025 0.006 7.310 40.35 25.84 22.48

35.11 22.48 0.00

Total = 156.15

d16=

d50= 0.125

d84= 0.195

130

Sample: HF_110809_59, 20cm, Unit1

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0.2 0.06 99.94

0.0197 0.500 0.999 0.33 0.09 99.85

0.0139 0.353 1.502 1.09 0.31 99.55

0.0098 0.249 2.006 6 1.68 97.86

0.0070 0.178 2.492 18.03 5.05 92.81

0.0049 0.124 3.006 16.27 4.56 88.25

0.0025 0.006 7.310 85.42 23.94 64.31

229.47 64.31 0.00

Total = 356.81

d16=

d50=

d84=

131

Sample: HF_110709_58, Bulk Top 10, SSS Unit 1

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0.05 0.01 99.99

0.0278 0.706 0.502 0.09 0.02 99.97

0.0197 0.500 0.999 2.25 0.55 99.41

0.0139 0.353 1.502 13.28 3.25 96.16

0.0098 0.249 2.006 30.8 7.54 88.62

0.0070 0.178 2.492 22.16 5.43 83.19

0.0049 0.124 3.006 18.99 4.65 78.54

0.0025 0.006 7.310 64.03 15.68 62.86

256.7 62.86 0.00

Total = 408.35

d16=

d50=

d84= 0.18

132

Sample: HF_110709_60

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0 0.00 100.00

0.0197 0.500 0.999 0.07 0.02 99.98

0.0139 0.353 1.502 0.25 0.06 99.93

0.0098 0.249 2.006 2.52 0.56 99.37

0.0070 0.178 2.492 25.6 5.71 93.66

0.0049 0.124 3.006 49.58 11.06 82.60

0.0025 0.006 7.310 150.25 33.50 49.10

220.21 49.10 0.00

Total = 448.48

d16=

d50=

d84= 0.125

133

Sample: HF_110709_61, Top 40, Unit 2

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0 0.00 100.00

0.0197 0.500 0.999 6.8 1.39 98.61

0.0139 0.353 1.502 22.47 4.60 94.01

0.0098 0.249 2.006 25.94 5.31 88.69

0.0070 0.178 2.492 35.88 7.35 81.34

0.0049 0.124 3.006 54.36 11.13 70.21

0.0025 0.006 7.310 261.96 53.65 16.56

80.87 16.56 0.00

Total = 488.28

d16=

d50=

d84= 0.2

134

Sample: HF_110709_62, Unit 3 , 20 cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0 0.00 100.00

0.0197 0.500 0.999 0.94 0.10 99.90

0.0139 0.353 1.502 4.2 0.46 99.44

0.0098 0.249 2.006 127.14 13.83 85.61

0.0070 0.178 2.492 101.74 11.07 74.53

0.0049 0.124 3.006 153.08 16.66 57.88

0.0025 0.006 7.310 349.09 37.99 19.89

182.79 19.89 0.00

Total = 918.98

d16=

d50=

d84= 0.24

135

Sample: HF_110709_63, Unit 3, 20cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0 0.00 100.00

0.0197 0.500 0.999 2.75 0.33 99.67

0.0139 0.353 1.502 4.03 0.49 99.18

0.0098 0.249 2.006 106.17 12.85 86.33

0.0070 0.178 2.492 82.46 9.98 76.36

0.0049 0.124 3.006 119.48 14.46 61.90

0.0025 0.006 7.310 424.74 51.39 10.51

86.86 10.51 0.00

Total = 826.49

d16=

d50=

d84= 0.24

136

Sample: HF_110709_64, Unit 3, 30cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0.07 0.01 99.99

0.0278 0.706 0.502 0.03 0.00 99.99

0.0197 0.500 0.999 15.84 2.24 97.75

0.0139 0.353 1.502 126.76 17.92 79.83

0.0098 0.249 2.006 65.81 9.30 70.53

0.0070 0.178 2.492 144.35 20.40 50.12

0.0049 0.124 3.006 157.61 22.28 27.84

0.0025 0.006 7.310 194.75 27.53 0.32

2.23 0.32 0.00

Total = 707.45

d16=

d50= 0.179

d84= 0.395

137

Sample: HF_110709_65, Unit 3, 40cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 2.92 0.51 99.49

0.0197 0.500 0.999 23.16 4.06 95.43

0.0139 0.353 1.502 89.11 15.61 79.81

0.0098 0.249 2.006 60.22 10.55 69.26

0.0070 0.178 2.492 126.88 22.23 47.03

0.0049 0.124 3.006 155.94 27.33 19.70

0.0025 0.006 7.310 109.61 19.21 0.50

2.83 0.50 0.00

Total = 570.67

d16=

d50= 0.19

d84= 0.39

138

Sample: HF_110709_66, Unit 3, 50cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 1.2 0.16 99.84

0.0197 0.500 0.999 19.55 2.63 97.21

0.0139 0.353 1.502 96.56 13.01 84.20

0.0098 0.249 2.006 85.15 11.47 72.73

0.0070 0.178 2.492 77.55 10.45 62.29

0.0049 0.124 3.006 82.09 11.06 51.23

0.0025 0.006 7.310 325.92 43.90 7.33

54.44 7.33 0.00

Total = 742.46

d16=

d50= 0.124

d84= 0.353

139

Sample: HF_110709_67, Unit 3, 60cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 2.69 0.57 99.43

0.0197 0.500 0.999 9.05 1.91 97.52

0.0139 0.353 1.502 74.94 15.80 81.72

0.0098 0.249 2.006 56.78 11.97 69.75

0.0070 0.178 2.492 123.26 25.99 43.75

0.0049 0.124 3.006 127.27 26.84 16.91

0.0025 0.006 7.310 76.25 16.08 0.83

3.95 0.83 0.00

Total = 474.19

d16= 0.124

d50= 0.195

d84= 0.355

140

Sample: HF_110709_68, Unit 3, 70cm

Size (in) Size (mm) Phi Mass (g) % Total % Finer

0.6250 15.875 -3.989 0 0.00 100.00

0.4380 11.125 -3.476 0 0.00 100.00

0.3120 7.925 -2.986 0 0.00 100.00

0.2230 5.664 -2.502 0 0.00 100.00

0.1570 3.988 -1.996 0 0.00 100.00

0.1100 2.794 -1.482 0 0.00 100.00

0.0787 1.999 -0.999 0 0.00 100.00

0.0555 1.410 -0.495 0 0.00 100.00

0.0394 1.001 -0.001 0 0.00 100.00

0.0278 0.706 0.502 0 0.00 100.00

0.0197 0.500 0.999 0.32 0.05 99.95

0.0139 0.353 1.502 99.86 14.99 84.96

0.0098 0.249 2.006 79.43 11.92 73.04

0.0070 0.178 2.492 79.08 11.87 61.17

0.0049 0.124 3.006 131.05 19.67 41.49

0.0025 0.006 7.310 249.16 37.40 4.09

27.23 4.09 0.00

Total = 666.13

d16=

d50= 0.15

d84= 0.353

141 8.3 Appendix C: Graphs of grain size (mm) vs. % finer

% Finer (3/1/09 HF_01)

70.00 60.00 50.00 40.00 % Finer 30.00 20.00 10.00 0.00 0.100 1.000 10.000 100.000

% Finer ( 3 /1 /0 9 HF_02)

120.00 100.00 80.00 60.00 % Finer 40.00 20.00 0.00 0.100 1.000 10.000 100.000

142

% Finer ( 3 /1 /0 9 HF_03)

120.00 100.00 80.00 60.00 % Finer 40.00 20.00 0.00 0.100 1.000 10.000 100.000

% Finer (HF_09_001)

120.00 100.00 80.00 60.00 % Finer 40.00 20.00 0.00 0.100 1.000 10.000 100.000

143

% Finer (HF_09_002)

120.00 100.00 80.00 60.00 % Finer 40.00 20.00 0.00 0.100 1.000 10.000 100.000

% Finer (HF_09_003)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

144

% Finer (HF_09_004)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_09_005)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

145

% Finer (HF_09_006)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_09_007)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

146

% Finer (HF_09_008)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_09_009)

80.00 70.00 60.00 50.00 40.00 % Finer 30.00 20.00 10.00 0.00 0.100 1.000 10.000 100.000

147

% Finer (HF_09_010)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110809_59)

102.00 100.00 98.00 96.00 94.00 % Finer 92.00 90.00 88.00 86.00 0.100 1.000 10.000 100.000

148

% Finer (HF_110709_58)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110709_60)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

149

% Finer (HF_110709_61)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110709_62)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

150

% Finer (HF_110709_63)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110709_64)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

151

% Finer (HF_110709_65)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110709_66)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

152

% Finer (HF_110709_67)

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

% Finer (HF_110709_68

120.00

100.00

80.00

60.00 % Finer

40.00

20.00

0.00 0.100 1.000 10.000 100.000

153 8.4 Appendix D: Sieving Statistics

d16 d16 (mm) d50 d50 (mm) d84

(mm) (w/o HF_01) (mm) (w/o HF_01) (mm)

Mean 0.4393 0.3611 1.2432 0.6619 1.5336

Standard Error 0.1405 0.1278 0.6005 0.1597 0.4723

Median 0.2375 0.2 0.5825 0.565 0.395

Mode #N/A #N/A 0.6 0.6 0.24

Standard Deviation 0.4866 0.4239 2.5475 0.6584 2.1645

Sample Variance 0.2368 0.1797 6.4900 0.4335 4.6851

Kurtosis 2.7177 9.2758 15.3904 3.2361 4.7530

Skewness 1.9527 2.9723 3.8300 1.8251 2.2830

Range 1.476 1.476 11.001 2.376 7.875

Minimum 0.124 0.124 0.124 0.124 0.125

Maximum 1.6 1.6 11.125 2.5 8

Sum 5.272 3.972 22.378 11.253 32.206

Count 12 11 18 17 21 Confidence Level 0.3092 0.2848 1.2669 0.3385 0.9853 (95.0%)

154 8.5 Appendix E: Wolman Count Statistics

a b

Mean 5.6048 3.8306 Standard Error 0.4156 0.2936 Median 5.5 4 Mode 4 4 Standard Deviation 2.3137 1.6348 Sample Variance 5.3532 2.6724 Kurtosis -0.7888 -1.0340 Skewness 0.1060 -0.1196 Range 8.5 5.5 Minimum 1.5 1 Maximum 10 6.5 Sum 173.75 118.75 Count 31 31 Confidence Level (95.0%) 0.8487 0.5996