<<

A Sequence Stratigraphic Analysis of the Allegheny Group (Middle ),

Southeast

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

Dreadnaught G. Stubbs

May 2018

© 2018 Dreadnaught G. Stubbs. All Rights Reserved. 2

This thesis titled

A Sequence Stratigraphic Analysis of the Allegheny Group (Middle Pennsylvanian),

Southeast Ohio

by

DREADNAUGHT G. STUBBS

has been approved for

the Department of Geological Sciences

and the College of Arts and Sciences by

Gregory C. Nadon

Associate Professor of Geological Sciences

Robert Frank

Dean, College of Arts and Sciences 3

ABSTRACT

STUBBS, DREADNAUGHT G. M.S., May 2018, Geological Sciences

A Sequence Stratigraphic Analysis of the Allegheny Group (Middle Pennsylvanian),

Southeast Ohio (253 words)

Director of Thesis: Gregory C. Nadon

This study examined a complete section of the Allegheny Group near Nelsonville in southeastern Ohio. The study site is located in the backbulge region of the

Alleghanian Orogeny where total accommodation was low. Facies analysis of six detailed measured sections concluded that the Allegheny Group was deposited by a fluvially dominated delta , which can be divided into ten sequences that vary from

2 to 15 m in thickness. Sequence boundaries were placed at the top of paleosol profiles and base of incised fluvial . Fluvial incision that occurred during the Falling

Stage and Lowstand Systems Tracts resulted in the erosion of some, or all, of an underlying sequence, the formation of paleosols on interfluves, and the amalgamation of some fluvial sandstones. Transgressive Systems Tract deposits include , marine and lacustrine prodelta mudstones, and fluvial sandstones. Highstand Systems Tract deposits consist primarily of interbedded laminated mudstones, siltstones. Vertical grain size pattern and fossil content were used to infer locations of the Maximum Flooding

Surfaces. The composition of individual sequences varies from a single fluvial body to a complex assemblage of marine to lacustrine mudstones, sandstones, and . The variation in facies patterns and sequence thickness within the section, and between this location and other locales along strike reported in the literature, is interpreted to be primarily a function of allogenic glacial-eustatic sea level changes. The allogenic driver was modified to varying degrees by autogenic processes, 4

such as channel avulsion during deposition of the Highstand Systems Tract, that shifted channel locations between sequences. 5

ACKNOWLEDGMENTS

My most genuine thanks to Dr. Greg Nadon for his supreme generosity in the form of wisdom, mentorship and kindness. I am honored and proud to have worked with you on these projects. Many thanks to my committee members Dr. Dan Hembree, Dr.

Dina Lopez and Dr. Craig Grimes for their patience and for sharing their expertise.

Thanks to the Ohio Geological Survey and especially Frank Fugitt. The field work in this study would have taken twice as long if it weren’t for Frank who was in the field with me every humid, Summer day. Thanks to Trey Hedrick who assisted me in the field.

Massive thank you to Cheri Sheets who helped me on almost a daily basis during my 4 in the department. I owe countless thanks to the faculty and staff of the Dept. of

Geol. Sciences at OU. I hope to stay in touch with you all.

Lastly, I would like to thank my family. I would not be in the position I am today if it wasn’t for your unwavering support.

Funding for this research is greatly appreciated and was provided by the Ohio

Department of Natural Resources ‘Ohio Rocks’ Grant and the Ohio University

Department of Geological Sciences Graduate Alumni Research Grant.

6

TABLE OF CONTENTS

Page

Abstract ...... 3 Acknowledgments ...... 5 List of Tables ...... 8 List of Figures ...... 9 CHAPTER 1 ...... 11 INTRODUCTION ...... 11 CHAPTER 2 ...... 16 PREVIOUS WORK ...... 16 CHAPTER 3 ...... 26 METHODOLOGY ...... 26 3.1 Introduction ...... 26 3.2 Measured Sections ...... 26 3.3 Petrography ...... 26 3.3 X-Ray Fluorescence (XRF) Analysis ...... 27 CHAPTER 4: ...... 29 LITHOFACIES AND FACIES ASSOCIATIONS ...... 29 4.1 Introduction ...... 29 4.2 Lithofacies ...... 29 4.2.1 Facies 1: Black Shale ...... 32 4.2.2 Facies 2: Mudstone ...... 35 4.2.3 Facies 3: Siltstone ...... 64 4.2.4 Facies 4: Sandstone ...... 64 4.2.5 Facies 5: Coal ...... 74 4.3: Facies Associations ...... 76 4.3.2 Facies Association 2: Subaqueous Delta ...... 79 4.3.3 Facies Association 3: Delta Plain Interfluve ...... 83 4.4 Fluvial Channel Deposits ...... 85 4.4.1 FA-4a Single Story Sandstone ...... 85 4.4.2 FA-4b Multistory Sandstone ...... 86 4.4.3 FA-4c Abandoned Channel ...... 87 4.5 Depositional Model – Fluvially Dominated Delta ...... 89 CHAPTER 5 ...... 94 SEQUENCE STRATIGRAPHIC FRAMEWORK ...... 94 7

5.1 Introduction ...... 94 5.2 Sequence Stratigraphy Terminology ...... 94 CHAPTER 6 ...... 104 CONCLUSIONS ...... 104 REFERENCES ...... 107 APPENDIX 1: DETAILED MEASURED SECTIONS ...... 117 APPENDIX 2: PALEOCURRENT DIRECTIONS ...... 123 APPENDIX 3: POINT COUNTING ...... 124 APPENDIX 4: XRD ...... 125 XRD Sample Preparation ...... 125 Brookville XRD Data ...... 127 Lower Kittanning XRD Data ...... 128 Upper Freeport Paleosol XRD Data ...... 129 APPENDIX 5: XRF DATA ...... 130 XRF Data Normalized to Total Mass ...... 132 XRF Standards ...... 133 APPENDIX 6: CIA AND MAP CALCULATIONS ...... 135 CIA-K and MAP Error Bar Calculations ...... 138 Upper Freeport ...... 138 Lower Kittanning ...... 139 Brookville ...... 141

8

LIST OF TABLES

Page

Table 4.1 Facies Table ...... 31 Table 4.2 Point Counting Results ...... 69 Table 4.3 Facies association table ...... 77

9

LIST OF FIGURES

Page

Figure 1.1. The Allegheny Group generalized section ...... 14 Figure 1.2. Study area ...... 15 Figure 2.1. Paleogeography ...... 22 Figure 2.2. Stratigraphic interval and sea level ...... 23 Figure 2.3. Marine incursion correlation ...... 24 Figure 2.4. Generalized cyclothem successions ...... 25 Figure 4.1. Measured section correllation ...... 30 Figure 4.2. Pyritized gastropod...... 34 Figure 4.3. Conchostrican ...... 34 Figure 4.4a. Sharp contact of black, fossiliferous mudstone ...... 43 Figure 4.4b. Erosional crossbedded sandstone ...... 44 Figure 4.5. Passively infilled trace ...... 45 Figure 4.6. Marine fossil assemblage ...... 46 Figure 4.8. Calcite cemented concretion ...... 47 Figure 4.9. Spherical concretions ...... 48 Figure 4.10. Rhizolith...... 49 Figure 4.11. Coalified plant matter ...... 50 Figure 4.12. Nodules with metallic luster...... 50 Figure 4.13. Mineral abundance from XRD analysis (UF)...... 51 Figure 4.14. Mineral abundance from XRD analysis (LK) ...... 52 Figure 4.15. Mineral abundance from XRD analysis (Brookville) ...... 53 Figure 4.16. Major element content from XRF analysis (UF) ...... 54 Figure 4.17. Major element content from XRF analysis (LK) ...... 55 Figure 4.18. Major element content from XRF analysis (Brookville) ...... 56 Figure 4.19. Geochemical indices (UF) ...... 57 Figure 4.20. Geochemical indices (LK) ...... 58 Figure 4.21. Geochemical indices (Brookville) ...... 69 Figure 4.22. CIA-K results with error ...... 60 Figure 4.23. MAP results on generalized stratigraphic section ...... 61 Figure 4.24. Gleyed LK paleosol ...... 62 Figure 4.25. MAP bar chart ...... 63 Figure 4.26. Erosional trough cross-bedded sandstone...... 68 Figure 4.27. Sandstone Petrography...... 70 10

Figure 4.28. Paleocurrent Results ...... 71 Figure 4.29. Lateral accretion surfaces ...... 72 Figure 4.30. Planar tabular crosbeddeding ...... 73 Figure 4.31. Middle Kittanning coal ...... 75 Figure 4.32. Appalachian Basin cyclothem model...... 82 Figure 4.33. Raft composed of coal and mudstone ...... 90 Figure 4.34. LPMS circulation ...... 91 Figure 4.35. Atchafalaya River drainage system...... 92 Figure 4.36. Paleocurrent/Paleogeography ...... 93 Figure 5.1. Parasequence set stacking pattern ...... 101 Figure 5.2. Sequence stratigraphic framework ...... 102 Figure 5.3. Sequence stratigraphic interpretations ...... 103

11

CHAPTER 1

INTRODUCTION

The Appalachian Basin is one of a number Late Paleozoic depocenters that contain extensive accumulations of sediment deposited during the Pennsylvanian

( - ; 320 - 298 Ma; Heckel, 2003). The thickness and sediment composition within each basin vary, but each was influenced by glacial-eustatic sea level changes of different temporal scales that alternately exposed and flooded the land surface (Heckel, 2008; Miall, 2016). The sedimentary response to those sea level changes was also influenced by the proximity to major siliciclastic source areas, such as the , and varying tectonic subsidence of the basement under the basins (Greb and Chesnut, 1996). The result is a of sediment packages that contain facies patterns, termed cyclothems by Wanless and Weller (1932), that were mapped across and between basins.

Previous studies of the Pennsylvanian-age succession in the Appalachian Basin correlated sections based on a combination of laterally extensive coal seams and marine bands (e.g., Sturgeon 1958). The coal deposits were the economic impetus for the basin studies over the past 150 years (Kosanke, 1988) and pollen provided the first means of rigorously correlating the coal beds within and between the Late Paleozoic basins. Those initial correlations were then corroborated and substantially augmented by of the marine bands (e.g., Boardman et al., 2004). The improved stratigraphic framework allowed the extension of the more complete marine sections of Mid-continent and basins into the siliciclastic-dominated Illinois and

Appalachian basins. 12

The stratigraphic section in the Appalachian Basin varies significantly between the regions adjacent to the mountains where total subsidence was relatively high and the distal portions of the basin, such as Ohio, where the section is thinned (Greb and

Chesnut, 1996). The Pennsylvanian series was initially subdivided based on the presence or absence of mineable coals. The section in Ohio consists of four major units variously termed formations or groups (Figure 1.1; Chapter 2). The Allegheny Group

(Moscovian-; ca. 309-305 Ma; Eble, 2002; Menning et al., 2006; Cecil, 2013) contained the most significant coal resources in Ohio and was extensively mapped at the county scale (Sturgeon, 1958). For example, Sturgeon (1958) compiled data from small roadcuts, streams, and mines to identify and name 13 cyclothems within the

Allegheny Group of Athens County. Those data were used by later workers to interpret the regional depositional environment of the Allegheny Group as a fluvially dominated delta analogous to the modern Mississippi Delta (e.g., Ferm, 1978). Nevertheless, correlation between sections measured by Sturgeon, even within townships, was hampered by the lack of detailed biostratigraphy and the accuracy of the topographic base maps that were available.

The improved understanding of the impacts of climate, tectonics, and glacial eustasy combined with the increased precision in stratigraphic correlation are now used to evaluate the local and regional controls on sediment accumulation of the Late

Paleozoic sections by combining process sedimentology with sequence stratigraphy

(Martino, 2016). This study uses sequence stratigraphic principles to reconstruct the differing contributions of allogenic and autogenic controls on the depositional history of the Allegheny Group.

Recent highway construction north of the town of Nelsonville in Athens and

Hocking Counties, Ohio (Figure 1.2) provides an ideal location to examine the vertical 13

variations in the Allegheny Group at a single location. The new road cuts expose the entire Allegheny Group along exit ramps of highway US 33 from the basal Homewood

Sandstone (uppermost Pottsville Group) to the Upper Freeport (No. 7) coal. The closely spaced sections provide a unique, three-dimensional exposure and the first opportunity to construct a sequence stratigraphic framework for the entire Allegheny Group in one location in this part of Ohio.

This study evaluated three hypotheses. The first hypothesis was that the main control on sediment deposition was 4th order glacio-eustatic changes is sea level. The second hypothesis was that the anomalously thick, white clay bed below the Lower

Kittanning (LK) coal is a paleosol that formed as a result of a prolonged sea level low- stand. The third hypothesis was that the Homewood Sandstone at the base of the section is the uppermost lithosome of the Pottsville Group rather than the basal sandstone of the Allegheny Group.

The results of the facies analysis showed that the fluvially dominated delta model

(e.g., Ferm, 1978) is still valid. Comparisons with regional and interbasinal stratigraphic compilations show that whereas climate varied somewhat during deposition of the

Group, the main controls on sediment accumulation in the study area were a combination of local variations in both the rate of change in relative rates of formation of accommodation and the total accommodation. The thick, light-colored bed beneath the

Lower Kittanning Coal is composed of two stacked paleosols. The higher concentration of light colored clay within the basal sandstone and the identification of the Putnam Hill

Shale above are considered sufficient to place the Homewood Sandstone in the

Pottsville Group, despite the fact that no petrographic difference was found in the framework grains between the Homewood Sandstone and the overlying Allegheny sandstone deposits. 14

Figure 1.1: The Allegheny Group forms the oldest significant deposit of Pennsylvanian age coal beds in Athens County (Sturgeon,1958). Average thickness for each lithology shown. PHS = Putnam Hill Shale, LK = Lower Kittanning Coal, MK = Middle Kittanning Coal, UF = Upper Freeport Coal.

15

Figure 1.2: Study area is under one square mile, located in Athens and Hocking Counties, Ohio. Highlighted in yellow are the six outcrops (OC-1 through 6) studied in this work.

16

CHAPTER 2

PREVIOUS WORK

2.1 Introduction

The study area is in southeastern Ohio on the north-west margin of the

Appalachian basin (Figure 2.1). The Allegheny Group was deposited during the Middle

Pennsylvanian (Heckel, 1989) when North America was rotated approximately 90° clockwise relative to the present position (e.g. Scotese, 2003; Blakey, 2011). The paleolatitude was between 7° and 10° south (Opdyke and Divenere, 1994). The possible controls on deposition were the amount of subsidence of the basin, the changes in eustatic sea level, and the climate. Each aspect has been studied extensively and will be briefly reviewed below prior to discussing the Allegheny Group deposits.

2.2 Tectonics

The Appalachian foreland basin was formed by thrust-loading of the North

American plate during the Alleghenian orogenic phase and the closure of the Rheic

Ocean leading to the formation of Pangea (Figure 2.1; Quinlan and Beaumont, 1984;

DeCelles and Giles, 1996). The space available for sediment accumulation in the backbulge zone of a foreland basin depends on the rheology of the bending plate and the amount of load applied at the orogen (Quinlan and Beaumont, 1984). Decelles and

Giles (1996) suggested that the elevation of the forebulge, absolute sea level elevation, and variability of sediment supply may be the important factors controlling deposition in the backbulge zone. Martino (2016) suggested that the stratigraphic repetition present 17

in the Pennsylvanian was in part a result of general slow subsidence that was somewhat accelerated at times.

The tectonic subsidence component in a backbulge region is expected to create space for sediment accumulation that will be similar over a broad area but with rates that can vary over time scales of 10 ky – 1 my (Plafker and Savage, 1970; Watts et al.,1982).

However, the thickness of backbulge deposits in the Late Paleozoic of the Appalachian basin are an order of magnitude greater than what would be expected if accommodation was simply due to plate flexure (Decelles and Giles, 1996) suggesting at least some of the subsidence was due to longer wavelength downwarping due to dynamic loading

(Decelles et al., 2011).

2.3 Paleoclimate

Between 320 and 300 Ma the Appalachian Basin and the adjacent Alleghenian highlands were largely situated in the tropical, equatorial belt (Figure 2.1; Scotese,

2003), characterized by long-term, humid to perhumid climate (Cecil et al., 2003; 2004).

The Appalachian basin was an east-west oriented depocenter located to the north of the rising orogeny at a paleolatitude of approximately 10 S at 320 Ma and 7 at 300 Ma

(Opdyke and Divenere, 1994). Palynological studies of the coals seams (Kosanke,

1988) suggested that the flora did not change significantly within the Allegheny Group implying that the climate was relatively consistent. This climate regime influenced the development of paleosol profiles.

The imprint of this climate regime influenced the development of paleosol profiles as the soil was in direct contact with the atmosphere when subaerially exposed (Driese and Ober, 2005; Hembree and Nadon, 2011). Cecil et al. (1985), Cecil (1990), and

DiMichele et al. (2010) suggested that the Appalachian basin climate shifted from 18

tropical and humid during deposition of the Allegheny Group to strongly seasonal with extended dry periods recorded within the (Upper Pennsylvanian).

Cecil et al. (2011) postulated that this shift was caused by a tectonically induced rain shadow effect from the growing central Pangean mountains after the initial formation of

Pangea.

2.4 Eustasy

Deposition in basins during the Middle Pennsylvanian was affected by global sea level changes on several time scales. The overall sea level rise that is present on the

3rd-order curve (Figure 2.2; Fischer, 1984; Ross and Ross, 1987; Dennison, 1989) was overprinted by high magnitude, 4th-order glacio-eustatic fluctuations (Heckel et al.,

2008). The Mid-continent basin and those farther to the west and south contain a more complete record of the transgressions and regressions during the Middle Pennsylvanian

(Boardman et al., 2004). Refinements to conodont zonation in the Pennsylvanian have allowed detailed correlations to be made between basins of the deposits of the major transgressions (Figure 2.3). Heckel concluded that the decrease of marine units within the Appalachian basin compared to those farther west was due to the proximity of sediment source of the Appalachian Mountains and the elevated position on the shelf of the LPMS (Figure 2.4).

2.5 The Allegheny Group

The term Allegheny Group was originally designated the Lower Productive

Measures because of the mineable coals present (Nadon and Hembree, 2007). The

Allegheny Group contains six marine members, which from base to top are the Putnam

Hill , Zaleski Flint, Vanport Limestone and Flint, Hamden Limestone and 19

Shale, the Washingtonville Shale, and the Door Run Shale. Sturgeon and Merrill (1949) concluded that the Dorr Run shale represented more restricted near-shore, estuary or lagoon environments. Most of the other marine members of the Allegheny group within

Ohio were interpreted as deposits of more distal settings because of the common occurrence of brachiopods (Sturgeon and Merrill, 1949). The major coal seams are named the Brookville, Clarion, Lower and Middle Kittanning, and the Lower and Upper

Freeport.

The Allegheny Group contains vertical repetitions of marine and terrestrial sediments successions typical of the Pennsylvanian strata in North America that were termed cyclothems by Wanless and Weller (1932). Sturgeon (1958) recognized 13 cyclothems containing variable amounts of marine and terrestrial sediments within the

Allegheny Group in Athens County (Figure 1.1). There was no single locality where all

Allegheny cyclothems and members were recognized and several units were present only in local areas. Stout (1939, 1947) considered the Middle Kittanning cyclothem to be the only complete example within the Allegheny Group.

The interest in coals within the Allegheny also brought attention to the laterally continuous, argillaceous units that underlie most coal beds termed underclay (Logan, 1842; Wanless, 1931; Hughes et al., 1992). Udden (1912) interpreted the underclays of the Pennsylvanian strata in Illinois as paleosols and suggested that the paleosols formed the top of each cycle whereas the base of the overlying coal bed is the base of the next cycle. According to Udden, a rise in the water table and the accumulation of plant matter (peat) initiated the next cycle. The contact between the underclay and the coal bed was, therefore, an unconformity.

Stout (1923, in, Cecil, 2003) attributed the origin of the Pennsylvanian underclays in Ohio to deposition in shallow basins in which plant life flourished but under conditions 20

which favored decay and not preservation of organic matter. Stout suggested that underclay formation was dependent on the duration of decomposition or preservation of organic matter and interpreted the underclays as having formed by subaqueous chemical alteration of sediment that was deposited in shallow-water marsh or swamp environments prior to the onset of peat formation.

The underclay below the Middle Kittanning coal in was studied

Gardner et al. (1988) who concluded that the unit represented a paleosol that was formed under a high water table. The paleosol beneath the Middle Kittanning coal in

Nelsonville was described and interpreted by Hembree (in Nadon and Hembree, 2007).

He concluded that the paleosol formed in very poorly drained conditions with the water table only a tens of centimeters below the surface. Evidence of a high water table included the gray color, preserved organic matter, small size and shallow depth of rhizoliths, and the absence of animal burrows. Based on the classification in Mack et al.

(1993) the paleosol below the Middle Kittanning coal was identified as a Gleysol.

The overall depositional system of the Allegheny has long been considered deltaic (e.g., Ferm, 1970), but the improved understanding of the effects of eustatic sea level changes (Van Wagoner et al., 1990) has changed some of the details of the controls on the depositional systems. For example, Flores (1968, 1979) interpreted the thick sandstone termed the Lower Freeport to be a deposit of a major delta distributary channel, whereas Flemming (2003) interpreted the lithosome as the product of an incised valley that formed after sea level had fallen but was infilled during the subsequent transgression.

Most studies of the Allegheny Group reported above relied on scattered, and sometimes widely spaced exposures, with little control on the possible lateral or vertical 21

variations in facies. This study is designed to examine the entire Allegheny Group to determine the controls on deposition through time at a single locality.

22

Figure 2.1: The diagram illustrates a cross-section across a typical foreland basin system (DeCelles and Giles, 1996). The distal backbulge portion where the study area is situated is highlighted by the yellow circle and red star. The bottom figure shows the location of the study area ~310ma during a highstand of sea level (Blakey, 2011). During this time there was a shallow inland sea termed the Late Pennsylvanian Midcontinent Sea (LPMS).

23

Figure 2.2: The stratigraphic position of the Allegheny Group relative to 3rd and 4th- order sea level fluctuations found in deposits of the Mid-continent basin after Nadon and Hembree, 2007. Change in paleolatitude after Opdyke and DiVenere (1994).

24

Figure 2.3: Late Middle and early Late Pennsylvanian marine incursions of the North American Midcontinent Sea during deposition of the upper half of the Allegheny Group (highlighted in red). The decrease in number of marine incursions in the Illinois and Appalachian basin were interpreted to be due to an increasingly elevated position on the shelf of the LPMS (Heckel, 2008).

25

Figure 2.4: Generalized cyclothem successions for the Midland, Midcontinent and Illinois and Appalachian basins (Heckel 2008). The thinness or absence of marine units in the Appalachian cycles was interpreted to be a result of low accommodation.

26

CHAPTER 3

METHODOLOGY

3.1 Introduction

The data gathered for this study consist of detailed measured sections, petrographic analysis of thin sections, and geochemical analyses of hand samples. The means used for collecting the data are described below.

3.2 Measured Sections

The newly exposed outcrops at US 33 exit 182 were described in detailed measured sections which include bed lithology, thickness, fossil content, grain size, color, sedimentary structures, and paleocurrent measurements. For a detailed grain size analysis, one 200-400 gm sample of each siltstone, clay and shale bed was collected at 25 cm intervals. Hand samples of sandstones were collected at the base, middle and top of beds to define coarsening or fining upward trends for lithofacies analysis. Grain size was measured to the nearest 1/2  using a stereo microscope and standard grain size chart. Mineralogical and textural observations, such as sorting, roundness, mineralogical anomalies, fossil content were also noted. The measured sections are presented in Appendix 1 and paleocurrent measurements in Appendix 2.

3.3 Petrography

Five thin sections were made from the hand samples collected in five of the major sandstone units in the study interval. The samples were initially cut into blanks using slab saws at Ohio University’s Department of Geological Sciences and then sent to Spectrum Petrographics to be made into thin sections. Three of the five sections 27

were stained for K-Feldspar. Framework grain mineralogy was obtained by point counting. A total 600 points were counted in each of the five thin sections using the number frequency, line method (Galehouse, 1969). The data are presented in Appendix

3.

3.3 X-Ray Fluorescence (XRF) Analysis

Samples of blocky mudstones were analyzed to obtain bulk rock geochemistry using X-ray fluorescence (XRF) to determine the vertical variations in elements that can be used to interpret the type of paleosol present. The major element geochemistry

(Appendix 3) was used to estimate the degree of chemical weathering from calculating the chemical index of alteration minus potassium (CIA-K) from which Mean Annual

Precipitation (MAP) values were calculated (Sheldon and Tabor, 2009; Nordt and

Driese, 2010).

In order to evaluate how much modern weathering may have altered the section

100-200 gm samples of blocky mudstone were collected at the exposed surface of the outcrop (weathered sample) and 20 cm horizontally into the outcrop (fresh sample). The fresh samples were free of limonite stains and their pedogenic structures were well preserved, whereas the weathered samples had iron stains from meteoric water percolation and friable texture from modern weathering. All samples were pulverized into pea-sized pellets and dried in an oven for 6 hours at 110°C. Each sample was then crushed into fine powder using an agate mortar and pestle. The powder was mixed 7:1 with GF-50 flux (6.000 g flux: 0.8500 g sample) and melted to form a glass bead by an

XRF Scientific furnace. Analyses were conducted on the Rigaku Supermini200 at Ohio

University and using the Fundamental Parameters (FP) method and calibrated against

12 USGS rock standards. BHVO-2 (basalt) and SBC-1 (Brush Creek Shale) were used 28

as running standards during analysis. No drift was seen in the standard measurements.

The XRF results returned with accuracy and standard deviation values for each oxide based on comparisons with standards. Absolute accuracy values were used to create the error bars as they were larger than the standard deviation and would therefor make the interpretation of noise less likely. Error bars were calculated by adding the absolute accuracy to the oxide wt% then dividing by the molecular wt. of the molecule to get the molecular proportion plus error. The molecular proportions minus error was subsequently calculated and then value was inserted into the CIA-K function. The error is the difference between each pair of values. The data are presented in Appendix 4.

3.4 X-Ray Diffraction (XRD) Analyses

The remaining powders from the XRF analysis were shipped to K/T Geoscience

Inc. for X-Ray Diffraction (XRD) analysis to determine the mineralogy of the samples.

The lab analyzed the <4micron fraction and reported the results as weight percent. The results and analytical procedures are reported in Appendix 5. Detection limits were reported to be between 1 - 5%.

29

CHAPTER 4:

LITHOFACIES AND FACIES ASSOCIATIONS

4.1 Introduction

To understand the controls on the deposition of the Allegheny Group it is first necessary to determine what processes formed the deposits. This is accomplished by identifying the lithofacies, which for siliciclastics are sediments deposited under similar hydrodynamic conditions (Miall, 2013). The lithofacies can be grouped into a few natural assemblages or associations that reflect the depositional environments that were present. The vertical and lateral changes in depositional environment can then be interpreted in terms of changes in relative sea level (Vail et al., 1977), which provides a means to interpret the lithofacies in a sequence stratigraphic, i.e., a chronostratigraphic, context.

4.2 Lithofacies

The six detailed measured sections in this study area (Figure 4.1) include five main lithologies, which can be subdivided into 10 facies and subfacies based on composition, grain size, and sedimentary structures. All the sediments, except for coal, are siliciclastic and vary in grain size from shale to sandstone (Table 4.1). The lithofacies are described below in order of increasing average grain size. The descriptions include field and microscopic observations and are followed by their respective interpretations.

30

Figure 4.1: Six measured sections along recent highway exposures that span the interval from the top of the Pottsville Group to the base of the Conemaugh Group. Marker beds include the Putnam Hill Shale (PHS) and Middle Kittanning (MKC) Coal.

31

Dominant Lithology Facies # Lithofacies Shale 1 Black shale 2a Black, fossiliferous mudstone Mudstone 2b Gray, laminated mudstone 2c Blocky mudstone Siltstone 3 Laminated siltstone

4a Trough crossbedded sandstone 4b Planar tabular crossbedded Sandstone sandstone 4c Rippled sandstone 4d Horizontally bedded sandstone

Coal 5 Coal

Table 4.1: Dominant lithologies and lithofacies observed in the study area.

32

4.2.1 Facies 1: Black Shale

Description: Shale, which is defined as fissile sediment composed of >50% clay-size particles (Schieber, 2011), makes up 2% of the total thickness of the measured sections.

Facies 1 occurs as beds that are horizontally laminated, black in color, and vary from

0.25 - 1 m in thickness. The laminations vary from 0.1 - 1 cm in thickness with no recognizable vertical thickening or thinning patterns. Facies 1 overlies coal (Facies 5) with a gradational basal contact and grades into gray mudstone (Facies 2b) or Siltstone

(Facies 3) above. Facies 1 lacks bioturbation and concretions but does show limonite stains on the weathered lamination surfaces. The most common organic remains present are carbonized fragments. One conchostracan and one internal mold of a gastropod were recovered from the uppermost exposure of Facies 1 in OC-4 (Figure 4.2,

4.3).

Interpretation: Facies 1 is indicative of a low energy, low oxygen environment (Johnson,

2004). The general lack of body fossils and burrows suggests a stressed environment, which may be a result of low dissolved oxygen in the water column (Dyni, 2006). The plant matter found in this facies indicates either rapid burial or an environment low in dissolved oxygen in which the terrigenous organic material was preserved. The gradational lower contacts indicate a gradual shift in depositional setting. The coarsening upward of the beds in coarser grained facies indicates a gradual shoaling and an increase in energy.

The depositional environment could have varied from marine to fresh-water depending on the stratigraphic location. No unequivocal marine fossils were found, although they have been reported from the Lower Freeport interval by Flores (1979).

The fossil gastropod fossil no evidence that could link it to a specific depositional setting, 33

and conchostracans can occur in brackish or fresh water conditions (Kozur and Weems,

2010). Grain size, lack of ichnofossils and sparse fossil assemblage suggests deposition in a low energy, low dissolved oxygen setting.

34

1mm

Figure 4.2: The pyritized internal mold of an oblong gastropod on a bedding plane with limonite staining from the uppermost beds of Facies 1 in OC-4. This example was found in the shale above the Upper Freeport coal.

1mm

Figure 4.3: A mold of a conchostracan valve in Facies 1 of OC-4, within the same bed as the gastropod in Figure 4.2 (Stigall, pers comm., 2017).

35

4.2.2 Facies 2: Mudstone

Mudstone is defined as a lithified mixture of clay and silt with < 10% sand-sized framework grains (Folk et al., 1970). Mudstones comprise a total of 11% of the measured sections and are divided into three subfacies based on color, structure, and fossil content.

4.2.2.1 Facies 2a: Dark Gray, Fossiliferous Mudstone

Description: Facies 2a occurs near the base of OC-1 and OC-6 and accounts for 4% of the sections. The beds of Facies 2a are dark gray (N3) in color, can be up to 3 m thick and have sharp and generally planar lower contacts with Facies 2c (Blocky Mudstone).

The silt content of the mudstone generally increases upwards. The upper contacts are erosional and capped by Facies 4a (Trough Crossbedded Sandstone) creating highly variable thickness (Figure 4.4). Over a lateral distance of 21 m the thickness varies from

3 m to zero. Along the basal contact of this subfacies at OC1 there are several locations where there are simple vertical shafts and plug-shaped extensions of the mudstone into the underlying bed of Facies 2c that are laminated, 2 cm in height and 1 cm in diameter

(Figure 4.5). Brachiopods, gastropods, and bivalves are common, and the laminations less prominent, in the basal 10 cm of Facies 2a at section OC-6 (Figure 4.6). Above the basal zone Facies 2a is less friable and contains distinctive, discontinuous layers of concretions that are generally 5 x 7 cm in cross-section, weather grayish-orange (10YR

7/4), and do not react with acid. The concretions make up approximately 5% of the bed, show compaction effects in some of the surrounding beds, and occasionally have brachiopods and bivalves preserved on the outer surfaces (Figure 4.6).

36

Interpretation: Facies 2a was considered the lateral equivalent of the Putnam Hill

Limestone by Merrill (1950) and is interpreted to be a low energy, normal marine deposit. The shallow, simple vertical shaft and plug-shaped features on the lower contact of Facies 2a are interpreted to be trace fossils. The horizontal internal laminations indicate passive fill of an open structure and the traces are interpreted to belong to the Glossifungites ichnofacies (MacEachern et al., 1992). The basal contact, therefore, is interpreted to be a marine flooding surface. The abundant shell material in the basal 10 cm implies a lower sedimentation rate and consequently the presence of a condensed section that formed early in the transgression. The content of basal portion of

Facies 2a and the location above terrestrial deposits (Facies 2c) are characteristics shared with the ravinement beds of the Pennsylvanian (Liu and Gastaldo, 1992;

Catuneanu and Zecchin, 2013; Zecchin et al., 2017).

The dark gray color of Facies 2a is a result of a high concentration of organic matter, which the preserved body fossils indicate was a result of rapid burial rather than anoxia. The compaction around at least some of the concretions, which are interpreted to be composed of siderite, indicate formation during early diagenesis. Siderite forms in anoxic conditions when the concentration of sulphate is low (Gautier, 1982; Berner,

1971). The presence of the concretions can indicate rapid burial, an influx of fresh water, or both (Potsma, 1982; Baird et al., 1986).

4.2.2.2 Facies 2b: Horizontally Bedded Gray Mudstone

Description: Facies 2b is gray mudstone, which makes up 7% of the exposures. The average bed thickness for is 0.7 m with laminae that vary from 0.1 - 1 cm in thickness.

Facies 2a occurs in all the exposures except OC-1. Beds of Facies 2b have gradational lower contacts with Facies 1. The upper contacts vary from gradational with Facies 2c 37

(Blocky Mudstones) and Facies 3 (Siltstone) to sharp when overlain by Facies 4a

(Trough Cross-bedded Sandstone) as seen in OC-4 above the Lower Kittanning Coal.

Beds of this facies generally have a coarsening upwards profile.

Black organic fragments up to 1 cm wide are common within Facies 2b, but no invertebrate fossils were observed. Red to brown, spherical grains 0.5 - 1mm diameter, which react with dilute HCL, are a common feature in Facies 2b. These spheres are found in groups of 5-30 in most samples, in others they are randomly distributed within the mud matrix (Figure 4.9). At one location there is a 1.5 cm-thick bed of nodules 1 mm in diameter separating Facies 2b from the horizontally bedded sandstone (Facies 4d) above it.

Interpretation: The mudstones of Facies 2b are interpreted to be low-energy deposits in water with relatively normal dissolved oxygen content (Collinson, 1986). Where Facies

2b grades into or out of Facies 1, the changes in grain size and color suggest that both energy and oxygen content were fluctuating, possibly in response to events such as storms (Schieber, 2011) or as a result of fluctuations in hyperpycnal flows (Bhattacharya and MacEachern, 2009). Such flows occur as a result of density differences between the water in the basin and the fluid-sediment mixture entering the depositional area.

Such flows are common in both marine and lacustrine deltaic systems (Tye and

Coleman, 1989a; Bhattacharya and MacEachern, 2009). The overall coarsening upward profile of this facies is indicative of an increase in energy up section consistent with an overall shoaling (Van Wagoner and Bertram, 1995).

38

4.2.2.3 Facies 2c: Blocky Mudstone

Description: Facies 2c occurs as beds of mudstone that break into angular blocks

(Retallack, 1988) with slickensides up to 30 cm in length on the surfaces except for the horizons directly beneath the coals, which are platy. The blocky mudstone facies varies from 0.23 to 3.05 m in thickness (average of 0.96 m) and occurs in all the sections.

Color varies from very pale orange (10yr 8/2) to black (N1) directly below coal seams, but is most commonly medium light gray (N6). The color occurs in bands that vary from

0.1 to 0.7 m in thickness. The lithology in the blocky mudstone facies varies from pure clay to mudstone with <25% silt. The basal contacts are gradational from Facies 3 and

2b.

Purple colored, vertical, lenticular tube features with relatively higher chroma, yellow dendritic structures ~25 mm in diameter were observed as casts in this facies only at OC-6 (Figure 4.10). No rhizohaloes were identified. Black organic material occurs in all beds of Facies 2c except those directly below the fossiliferous mudstones of

Facies 2a (OC-1, OC-6) The organics vary in size from 1 mm in width and up to 10 cm long (Figure 4.11). The 0.5-1 mm in diameter spherical, red-brown grains observed in

Facies 2b are also common in the gray-colored beds of Facies 2c. Spherules of similar size to the red-brown spheres, but with a metallic luster, are also common in Facies 2c and stand out against the mud matrix when examined under a stereomicroscope (Figure

4.12).

The XRD analyses show that the most common clay mineral within Facies 2c is kaolinite, which can comprise as much as 88% of the clays in the < 4 micron fraction

(Fig. 4.13, 4.14 and 4.15 show variations in clay mineralogy; Appendix 4). The amount of kaolinite decreases with depth in the upper and lower profiles sampled. The bed below the Lower Kittanning Coal has two decreasing patterns, each approximately 0.75 39

m thick. The percentage of illite + mica and smectite tends to increase with depth in all three beds with a mid-profile break in the Lower Kittanning bed.

The XRF analyses show that the most abundant oxide was SiO2 66-71% in the basal bed and 60-72% below the Lower Kittanning Coal, and 61-62% below the Lower

Freeport Coal (Figure 4.16, 4.17, 4.18). As with the clay mineralogy, the Lower

Kittanning profile has break below 75 cm depth. The vertical trends in ratios of different oxides are used to determine characteristics of paleosols (Sheldon and Tabor 2009).

The ratios indicating hydrolysis (Bases/Al, Al/Si, Ba/Sr) show varying trends (Figure

4.19, 20, 4.21). The Al/Si all decrease with depth whereas the Bases/Al generally increase with depth. The Ba/Sr values are more widely scattered, but increase with depth in the lower two profiles and increase in the upper profile. Oxidation ratios

(Fetot/Al, (Fetot+Mg)/Al) both show slight increases with depth in all three profiles (Figure

4.19, 20, 4.21). Salinization ratios ((K+Na)/Al, Na/K, Na/Al) show increases with depth for all three profiles, but are more variable in the Lower Kittanning profile (Figure 4.17).

Base loss with depth (Ca, Mg, Na, K)/Ti) increases but, again, is more variable in the

Lower Kittanning profile.

The XRF data also allow a calculation of weathering using the CIA-K

(Al/(Al+Ca+Na) * 100) and CALMAG (Al/(Al+Ca+Mg) * 100) ratios. The values for CIA-K are consistently higher than those calculated using CALMAG (Figure 4.19, 4.20, 4.21).

All but those in the upper profile are > 93% (CIA-K). The CALMAG values for the upper profile vary from 88-91%. The weathering ratios are used to calculate mean annual precipitation (MAP) estimates (Figure 4.19, 4.20, 4.21). Those calculated using CIA-K are similar for the basal profiles but decrease at the level of the Lower Freeport Coal.

The CALMAG values have a similar trend but are from 226 to 369 mm/yr higher than 40

those values calculated using CIA-K. Results shows that, overall there was little difference between the surface and trenched samples (Figure 4.22).

Interpretation: The blocky mudstones are interpreted to be paleosols. Soils are complex stratigraphic units formed at the surface of the earth. The variability of modern soils is evident in the complexity of soil taxonomy (e.g., Boul et al., 1989). The recognition and interpretation of soil profiles in the stratigraphic record is made more challenging by the temporal variations in atmosphere composition, climate, biota, and diagenesis (Mack et al. 1993). The latter is particularly important because some of the criteria used to define modern soils are not preserved in the stratigraphic record. Mack et al. 1993 proposed a simplified nomenclature for soil orders, and the criteria that can be used to identify them, to account for the loss of information through time.

The colors, blocky fractures, and slickensides on the surfaces are indicators of paleosols (e.g., Retallack, 1988). The dendritic structures in section OC-6 are interpreted to be rhizoliths (Kraus and Hasiotis, 2006). The decrease in kaolinite with depth as the illite + micas value increase is interpreted to be a result of weathering of the upper 50 - 75 cm of the profiles. The type of paleosol present can be interpreted using the structures,horizonation, and clay mineral content.

The red/brown 1mm in diameter spheres are interpreted to be sphaerosiderite based on appearance in hand sample and thin section, and likeness to chemically proven siderite spherules (Unfar et al., 2004, Driese and Ober, 2005, Ludvigson et al.,

2013, Rosenau et al., 2013) in the blocky mudstones of the Late Pennsylvanian. No thin sections of the blocky mudstone were analyzed, however, identical spheres were within the sandstone thin sections. Siderite is a common mineral precipitate in poorly drained

‘‘wetland’’ soils and occurs as concretions, or spherules (Driese and Ober, 2005). The 41

possible presence of siderite helps define the paleohydrologic and geochemical environments associated with underclay saturation. Stability fields for siderite require the ground water have low sulfur activities, high partial pressure of carbon dioxide, alkaline conditions, and anoxic, Fe-reducing pore waters, (Gardner et al., 1988).

The dominance of kaolinite in the clay mineral suite is indicative of warm humid climate (Sheldon and Tabor, 2009). The presence of slickensides indicates shrinking and swelling of clays. Smectite is not the most abundant clay, however, smectite alters to illite with time and the combination of illite + mica and smectite is 21-36% of the UF paleosol, 6-23% of the LK paleosol and 12-32% of the Brookville paleosol (Appendix 4).

However, slickensides alone are not enough evidence to definitively claim a climate regime because they could develop due to compaction after burial (Retallack, 1990).

Sheldon and Tabor (2009) suggested that if the Bases/Al ratio was < 0.5 then the profile could be termed a eutric Argillisol (i.e., paleoUltisol), particularly if combined with a high kaolinite content which is common in modern Ultisols. All three profiles have ratios < 0.5 and all three have relatively high kaolinite content, however, the combination of shrink-swell structures, preserved organic matter, gray color, small rhizoliths, sphaerosiderite and gleyed, low chroma colors of the paleosols beneath the Lower

Kittanning and Upper Freeport coals in this section are interpreted to be the result of a rise in water table and subsequent flooding of paleoVertisols, similar to the interpretations of paleosols in the Early and Mid-Late Pennsylvanian of the Central

Appalachian Basin (Figure 4.24, Mack et al., 1993, Driese and Ober, 2005, Nadon and

Hembree, 2007, Martino, 2016). The basal profile, which also contains the most silica is not complete due to erosion on the transgression and might qualify as the lower portion of an eutric Argillisol. 42

Rosenau et al. (2013) characterized paleosols equivalent to those below the

Lower Kittanning in the Illinois Basin as gleyed Protosols and gleyed Vertisols. In the same study, the Illinois Basin equivalent to the Middle Kittanning paleosol is termed it a gleyed Vertisol and gleyed calcic Vertisol and the Upper Freeport paleosol in the Illinoian

Basin is termed a gleyed vertic Calcisol.

The presence of the discontinuity in the mineralogy and major element geochemistry in the Lower Kittanning profile is interpreted to indicate two, stacked paleosol profiles. Sturgeon (1958) suggested that the Lawrence coal and underclay, which is present in counties to the south (Stout, 1923), might be amalgamated with the

Lower Kittanning underclay in Athens County and these data are consistent with that interpretation.

The MAP values calculated from the weathering indices for the basal two profiles are similar to that reported for the Middle Kittanning paleosol by Nadon and Hembree

(2007). The decrease between the Lower Kittanning profile and the Upper Freeport profiles varies 2% for the CIA-K equation (Figure 4.25) to 9% using the CALMAG ratio.

This decrease is slight, but, consistent with the overall drying that occurred during deposition of the basal Conemaugh Group (e.g., Kosanke and Cecil, 1996; Hembree and Nadon 2011). This gradual, long term climate change is consistent with current model of a rain shadow caused by central Pangean orogenic events in the Late

Pennsylvanian (Cecil et al., 2011). The use of the CIA-K value is reasonable for the basal paleosol profile, however, if the upper profiles are paleoVertisols, and the gley overprinting did not affect the elemental ratios, then the CALMAG parameter is more appropriate since Nordt and Driese (2010) proposed using the CALMAG parameter for solely for Vertisols.

43

0.5 m

Figure 4.4a: Black fossiliferous mudstone with siderite concretions including gastropods, brachiopods, and bivalves. Facies 2a has a sharp upper contact with an erosional overlying sandstone in both outcrops where it is exposed (OC-1 and OC-6).

44

Figure 4.4b: The upper contact of Facies 2a mudstone eroded by a crossbedded sandstone at OC-6.

45

Figure 4.5: Black fossiliferous mudstone layer (Facies 2a) at OC-1. Along the sharp basal contact there is a wedge/bulb shaped trace interpreted to be of the Glossifungites ichnofacies extending into the underlying blocky mudstone bed (Facies 2c).

46

A.

B. C.

Figure 4.6: Facies 2a fossil assemblage, A.) Articulate brachiopod molds, B.) Bivalve along bedding plane, C.) Small (1 cm wide) articulate brachiopod observed on surface of concretion.

47

Figure 4.8: Calcite cemented Iron-carbonate concretion found in the horizontally bedded mudstone of Facies 2c interbedded with Facies 2a above the LK coal bed (OC-4).

48

10mm

m

1 mm

Figure 4.9: Spherical iron-carbonate concretions settled in gray mud matrix in a Facies 2c bed of OC-3. These spherical concretions are also observed in Facies 2b, 3 and 4.

49

1cm

Figure 4.10: Purple-gray vertical tubular trace observed in the blocky mudstone (Facies 2c) of OC-6, interpreted to be a rhizolith. Orange-yellow surfaces are iron stains caused by modern meteoric water percolation.

50

Figure 4.11: Coalified plant matter found in the blocky mudstone below the Middle Kittanning coal (OC-2)

1mm

Figure 4.12: Nodules with metallic luster in Facies 2c. These are similar in size to the sphaerosiderite and appear in a pure clay matrix. Observed in the blocky mudstones below the Upper Freeport and Lower Kittanning coal beds.

51

Upper Freeport Paleosol XRD Results Kaolinite (%) Illite & Mica (%)

40 50 60 70 80 90 0 10 20 30 40 0.0 0.0

25.0 25.0

50.0 50.0 Depth (cm) Depth (cm) 75.0 75.0

Quartz (%) Chlorite (%) 0 5 10 15 2 3 4 5 6 0.0 0.0

25.0 25.0

50.0 50.0 Depth (cm) Depth (cm) 75.0 75.0

K-Spar (%) Smectite (%) 0.0 0.5 1.0 1.5 2.0 0 1 2 3 4 0.0 0.0

25.0 25.0

50.0 50.0 Depth (cm) Depth (cm) 75.0 75.0

Kaolinite/Mica 0 5 10 15 20 0.0

25.0

50.0 Depth (cm) 75.0

Figure 4.13: Mineral abundance from XRD analysis of the Facies 2c Upper Freeport paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 4).

52

Lower Kittanning Paleosol XRD Results

Kaolinite (%) Illite & Mica (%) 40 50 60 70 80 90 0 10 20 30 40 0 0 50 50 100 100 150 150 Depth (cm) Depth (cm) 200 200

Quartz (%) Chlorite (%)

0 5 10 15 2 3 4 5 6 0 0 50 50 100 100 150 150 Depth (cm) Depth (cm) 200 200

K-Feldspar (%) Smectite (%)

0.0 0.5 1.0 1.5 2.0 0 1 2 3 4 0 0 50 50 100 100 150

Depth (cm) 150 Depth (cm) 200 200

Kaolinite/Mica

0 5 10 15 20 0 50 100

150 Depth (cm) 200

Figure 4.14: Mineral abundance from XRD analysis of the Facies 2c Lower Kittanning paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 4).

53

Brookville Paleosol XRD Results

Kaolinite (%) Illite & Mica (%) 40 50 60 70 80 90 0 10 20 30 40 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) 100 100

Quartz (%) Chlorite (%) 0 5 10 15 2 3 4 5 6 0 0 25 25 50 50

Depth (cm) 75 75 Depth (cm) 100 100

K-spar (%) Smectite (%) 0.0 0.5 1.0 1.5 2.0 0 1 2 3 4 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) 100 100

Kaolinte/Mica 0 5 10 15 20 0 25 50 75 Depth (cm) 100

Figure 4.15: Mineral abundance from XRD analysis of the Facies 2c Brookville paleosol in OC-6. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 4).

54

Upper Freeport Paleosol XRF Results

SiO2 (wt%) Al2O3(wt%) 50 60 70 80 20 25 30 35 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm)

K2O (wt%) Total Fe (wt%) 0 2 4 6 1 2 3 4 5 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) TiO2(wt%) MgO(wt%) 0.5 1.5 2.5 3.5 0.0 0.5 1.0 1.5 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) CaO (wt%) P2O5(wt%) 0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) MnO (wt%) Na2O(wt%) 0.00 0.02 0.04 0.06 0.08 0.0 0.1 0.2 0.3 0.4 0 0 25 25 50 50 75 75 Depth (cm) Depth (cm) Figure 4.16: Major element content from XRF analysis of the Facies 2c Upper Freeport paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 5).

55

Lower Kittanning Paleosol XRF Results Al O (wt%) SiO2 (wt%) 2 3 50 60 70 80 20 25 30 35 0 0 50 50 100 100 150 150 200 Depth (cm) Depth (cm) 200

K2O (wt%) Total Fe (wt%) 0 2 4 6 1 2 3 4 5 0 0 50 50 100 100

150 150 Depth (cm)

Depth (cm) 200 200

TiO2 (wt%) MgO(wt%)

0.5 1.5 2.5 3.5 0.0 0.5 1.0 1.5 0 0 50 50

100 100 150 150 200

Depth (cm) 200 Depth (cm) CaO(wt%) P2O5 (wt%)

0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3

0 0

50 50 100 100 150 150

200 Depth (cm) 200 Depth (cm) MnO (wt%) Na2O (wt%)

0.00 0.02 0.04 0.06 0.08 0.0 0.1 0.2 0.3 0.4

0 0 50 50 100 100 150 150

200 Depth (cm) Depth (cm) 200

Figure 4.17: Major element content from XRF analysis of the Facies 2c Lower Kittanning paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 5).

56

Brookville Paleosol XRF Results

SiO2(wt%) Al2O3 (wt%) 50 60 70 80 20 25 30 35 0 0 25 25 50 50 75 75 100 100 Depth (cm) Depth (cm) K O (wt%) 2 Total Fe (wt%) 0 2 4 6 1 2 3 4 5 0 0 25 25 50 50 75 75 100 100 Depth (cm) Depth (cm)

TiO2 (wt%) MgO (wt%) 0.5 1.5 2.5 3.5 0.0 0.5 1.0 1.5 0 0 25 25 50 50 75 75 100 100 Depth (cm) Depth (cm)

CaO (wt%) P2O5 (wt%) 0.0 0.2 0.4 0.6 0.0 0.1 0.2 0.3 0 0 25 25 50 50 75 75 100 100 Depth (cm) Depth (cm)

MnO (wt%) Na2O (wt%) 0.00 0.02 0.04 0.06 0.08 0.0 0.1 0.2 0.3 0.4 0 0 25 25 50 50 75 75 100 100 Depth (cm) Depth (cm) Figure 4.18: Major element content from XRF analysis of the Facies 2c Brookville paleosol at the top of the Pottsville Group (OC-6). Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep (see also Appendix 5).

57

Upper Freeport Paleosol Weathering Indices

Hydolysis 1 Hydolysis 2 0.0 0.2 0.4 0.0 4.0 8.0 12.0 0 0 Sum Bases/ Ba/Sr Al 25 Al/Si 25 Depth (cm) Depth (cm) 50 50

Oxidation Salinization 0.0 0.1 0.2 0.3 0.00 0.10 0.20 0 Fetot/Al 0 (K+Na)/ Al (FeTot+M 25 Na/K g)/Al 25 Depth (cm) Depth (cm) Na/Al 50 50

Provenance Base Loss 0.0 0.1 0.2 0.0 1.0 2.0 3.0 4.0 0 0 Ca/Ti Ti/Al Mg/Ti 25 25 Na/Ti Depth (cm) Depth (cm) 50 50 K/Ti

Weathering Index MAP (mm/) 85 90 95 100 1000 1500 2000 0 0 CIA-K CIA-K 25 CALMAG 25 CALMAG Depth (cm) Depth Depth (cm) 50 50

Figure 4.19: Geochemical indices from Retallack 2001, Sheldon and Tabor 2009, which utilize XRF data to calculate proxies including hydrolysis, oxidation, salinization, provenance, base loss, weathering intensity and mean annual precipitation (MAP) of the Facies 2c Upper Freeport paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep.

58

Lower Kittanning Paleosol Weathering Indices

Hydrolysis 1 Hydroysis 2 0.0 0.2 0.4 0.0 4.0 8.0 12.0 0 0 25 Sum 25 50 Bases/Al 50 Ba/Sr 75 75 100 Al/Si 100 125 125 Depth (cm) Depth (cm) 150 150 175 175

Oxidation Salinization 0.0 0.1 0.2 0.3 0.00 0.10 0.20 0 0 (K+Na)/Al 25 Fetot/Al 25 50 50 Na/K 75 75 100 (Fetot 100 Na/Al 125 +Mg)/Al 125 Depth (cm) Depth (cm) 150 150 175 175

Provenance Base Loss 0.0 0.1 0.2 0.0 1.0 2.0 3.0 4.0 0 0 Ca/Ti 25 25 50 Ti/Al 50 Mg/Ti 75 75 100 Na/Ti 125 100 150 125 K/Ti Depth (cm) 175 Depth (cm) 150 200 175

Weathering Incdex MAP (mm/year) 85 90 95 100 1000 1500 2000 0 0 25 CIA-K 25 CIA-K 50 50 75 CALMAG 75 CALMAG 100 100 125 125 Depth (cm) 150 Depth (cm) 150 175 175

Figure 4.20: Geochemical indices from Retallack 2001, Sheldon and Tabor 2009, which utilize XRF data to calculate proxies including hydrolysis, oxidation, salinization, provenance, base loss, weathering intensity and mean annual precipitation (MAP) of the Facies 2c Lower Kittanning paleosol in OC-4. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep.

59

Brookville Paleosol Weathering Indices

Hydolysis 1 Hydolysis 2 0.0 0.2 0.4 0.0 4.0 8.0 12.0 0 (Ca+Na 0 +K+Mg)/ 25 25 Al Ba… 50 Al/Si 50 Depth (cm) Depth (cm) 75 75

Oxidation Salinization 0.0 0.1 0.2 0.3 0.00 0.20 0 0 Fetot/Al (K+Na) 25 25 /Al 50 (Fetot+M 50 Na/K Depth (cm) g)/Al Depth (cm) 75 75

Provenance Base Loss 0.0 0.1 0.2 0.00 2.00 4.00 0 0 Ca/Ti Ti/Al 25 25 Mg/Ti 50 50 Na/Ti Depth (cm) Depth (cm) K/Ti 75 75

Weathering Index MAP (mm/year) 85 90 95 100 1000 1500 2000 0 0 CIA-K 25 CIA-K 25

50 CALMAG 50 CALMAG Depth (cm) Depth (cm) 75 75

Figure 4.21: Geochemical indices from Retallack 2001, Sheldon and Tabor 2009, which utilize XRF data to calculate proxies including hydrolysis, oxidation, salinization, provenance, base loss, weathering intensity and mean annual precipitation (MAP) of the Facies 2c Brookville paleosol in OC-6. Displayed are only the ‘Fresh’ samples collected from the trench 20 cm deep.

60

Figure 4.22: CIA-K results from the XRF analysis of three blocky mudstones below two coals and Facies 2a near the base of the section. The error bars only displayed for the fresh samples only. Of the 15 pairs, only two samples were appreciably different from the other or outside of calculated error bar (Appendix 6). There is a general decreasing trend of CIA-K values from the top to the base of the profiles in each bed of Facies 2c. This is a typical weathering pattern in soils and paleosols.

61

Figure 4.23: Results from the XRF analysis in alignment with the sampling location within the generalized measured section of the study area. Line plots are Mean Annual Precipitation (MAP) vs. depth. Fresh samples are represented by the blue lines and weathered samples by the red lines. A decreasing trend in weathering/precipitation is observed in each of the three sample sets when observing the fresh samples.

62

Figure 4.24: The photograph is of the Lower Kittanning coal bed (LK cb) and blocky mudstone of OC-4. It is interpreted that base level rise overprints paleo-Vertisols, causing the resulting gleyed color and siderite concretions to be present with shrink swelling structures, alike to the paleosols described in Driese and Ober, 2005 and Martino, 2016. When base level is high peat mire form and lead to coal deposits. As the sea inundates the peat mires fined grained siliciclastic sediments are deposited. CIA-K and MAP values from the paleoclimate analysis may be inflated if this interpretation is correct and the drowning led to further depletion of alkali and alkaline earth elements (Ca, Mg, Na) within the paleosol.

63

Average MAP from CIA-K (mm/year) Paleosol Fresh Upper Freeport Paleosol 1347.08 Lower Kittanning Paleosol 1369.31 Brookville Paleosol 1374.34

CIA-K Ave MAP (mm/year) 1600 1500 1400 1300 1200 1100 1000 Upper Freeport Lower Brookville Paleosol Kittanning Paleosol Paleosol

Average MAP from CALMAG (mm/year) Paleosol Fresh Upper Freeport Paleosol 1590.89 Lower Kittanning Paleosol 1723.21 Brookville Paleosol 1714.67

CALMAG Ave. MAP (mm/year) 1900 1800 1700 1600 1500 1400 1300 1200 Upper Freeport Lower Brookville Paleosol Kittanning Paleosol Paleosol

Figure 4.25: Bar charts and supplementary tables show the trends in Mean Annual Precipitation (MAP). Error bars for MAP calculated from the CIA-K and CALMAG weathering indices are ± 182 mm/year and ± 108 mm/year, respectively (Sheldon and Tabor 2009; Nordt and Driese, 2010).

64

4.2.3 Facies 3: Siltstone

Description: Laminated siltstone is observed in many stratigraphic levels and locations within the study area. Facies 3 makes up a total of 14% of the strata with bed thickness varying from 0.20 - 4.06 m (average =1.17 m). Both the basal and upper contacts are typically gradational except when the overlying facies is sandstone in which case the contact is erosional. The only structures present are horizontal laminations ≤ 1 cm thick.

The siltstones are medium gray except beneath Facies 2c at the base of the section of

OC-4 and OC-6 where it is lighter gray. Red-brown spheres less than 1 mm in diameter occur throughout this facies. Small (<1mm) black flakes of carbon and muscovite are abundant, whereas biotite is present, but not common.

Interpretation: The laminated siltstone facies was deposited under higher energy conditions than the mudstone facies. The horizontal laminations formed as traction deposits from events, such as hyperpyncnal flows (e.g., Bhattacharya and MacEachern,

2009). The small carbonized flakes are organic debris from a terrestrial setting. The spheres are interpreted as siderite and indicative of fresh water deposition or infiltration

(Potsma, 1982). The absence of bioturbation and invertebrate fossils suggests an environment stressed by changes in factors such as salinity, turbidity, or energy

(Pemberton and Wightman, 1992; Pearson et al., 2012).

4.2.4 Facies 4: Sandstone

Sandstone within the study interval makes up 58% of the sections measured.

The sandstones are divided into four subfacies (Table 4.1) based on the sedimentary structures within the beds.

65

4.2.4.1 Facies 4a: Trough Cross-bedded Sandstone

Description: Facies 4a represents 48% of the measured sections. The basal contacts are erosional with relief on the base that varies from a less than 0.1 to 5 m (Figure 4.26).

The beds of this facies are lower medium-grained at the base and upper very fine- grained sandstone at the top. Clay clasts and iron concretions are found in the basal 15-

20 cm of Facies 4a where the bottom contact has high relief. The upper contact of this subfacies is generally gradational into laminated siltstone (Facies 3) or horizontally bedded sandstone (Facies 4d). Facies 4a overlies a wide range of lithologies, but most commonly it lies above another trough cross-bedded sandstone. In the middle portion of the thick Facies 4a exposure in OC-1, there is a <35 cm thick siltstone lens with iron- carbonate concretions up to 22 cm in diameter. Cementation by calcite is common but variable. Carbonized organic material is present throughout this subfacies. Fragments up to 5 cm wide and 6 cm long, including lycopsid stem molds were observed.

The sandstone of Facies 4a is most commonly a lithic arenite (Table 4.2; Figure

4.27) with abundant muscovite and brown, red-weathering spheres like those found in

Facies 1, 2a, 2b, and 2c. Paleocurrent measurements from Facies 4a average 330° (n =

50) (Figure 2.28, Appendix 2). Internally, the trough crossbeds average 15 - 20 cm in thickness with a maximum of 60 cm. Reactivation surfaces locally truncate cross- bedding. At location OC-1 there are large epsilon cross strata within this facies (Figure

4.29) which are parallel to the basal erosional surface of the sand which cuts through both a Facies 2a and 2c unit. The epsilon cross strata are 3.25 m in height and approximately 28 m wide. The basal ~25cm of this bed contain abundance clay clasts and red/brown concretions resembling the concretions and clay deposits from the underlying Facies 2a.

66

Interpretation: The trough cross-beds of Facies 4a represent deposits of sinuous crested bedforms under unidirectional flow (Rubin and Carter, 1987). The clay clasts found in the lower portions of the bed are interpreted to be the remnants of Facies 2a beds eroded by high energy water currents. The epsilon crossbedding observed in OC-1 are interpreted to be drapes on the underlying erosion surface rather than lateral accretion surfaces from point bar migration.

4.2.4.2 Facies 4b: Planar-tabular Cross-bedded Sandstone

Description: Planar-tabular cross-beds were observed within the basal sandstone and near the top of the thick sandstone in OC-1 as well as in the upper portion of the sandstone body in OC-5. Facies 4b overlies either Facies 4a or 4d with a sharp contact

(Figure 4.30). The beds are composed of lower very fine-grained, lithic arenite (Figure

4.27, sample OC5-RLB1*) with sand-sized concretions scattered throughout. Bedforms vary from 15-30 cm in thickness. Grains are moderately well sorted and angular to subangular.

Interpretation: Planar tabular cross-stratification is the result of migrating straight-crested dunes in a unidirectional current and likely with lower velocity than Facies 4a (Rubin and

Carter, 1987).

4.2.4.3 Facies 4c: Rippled Cross-laminated Sandstone

Description: Sandstone with ripple cross lamination is present in two locations in with an average grain size of upper very fine-grained and relatively poor sorting. The deposits vary from 1 - 5 cm in thickness and occur between beds of laminated siltstone (Facies

3). The basal contacts are sharp and the top contacts gradational. The most notable 67

characteristic of this facies is the abundance of red-brown, sand-sized spheres that make the cross laminations easily visible.

Interpretation: Cross-laminations in Facies 4c were the result of lower flow regime unidirectional flow (Harms et al., 1982). The thin alternations with the siltstones suggests fluctuating current strength during flow events. The location of these beds within finer grained siltstones suggests deposition from more proximal hyperpycnal flows

(Bhattacharya and MacEachern, 2009)

4.2.4.4 Facies 4d: Horizontally Bedded Sandstone

Description: Horizontally bedded sandstone makes up 9% of the strata in the study area. Bed thickness varies from 0.35 - 3 m (Ave. 1.1 m). Internal laminations and bedding range from 0.2 to 2 cm in thickness. Basal contacts are sharp when Facies 4d overlies sandstones and the upper contacts are gradational into mudstones and siltstones. Grain size varies from lower very fine- to upper fine-grained. Grains are sub- angular and poorly to moderately well sorted. In hand sample, mineralogy appears similar to Facies 4a, i.e., mostly composed of quartz with abundant muscovite, red- brown sand-sized grains, and carbon flakes.

Interpretation: The horizontal lamination is interpreted as the product of upper flow regime plane bed (Allen, 1984). Fine grain sizes and shallow flow depths are capable of producing this structure (Allen, 1984). 68

A. B.

Figure 4.26: Figure A; Trough cross-bedded sandstone (Facies 4a) cutting through the black, fossiliferous, marine mudstone (Facies 2a) and blocky mudstone (Facies 2c) in OC-1. Figure B; The same photo as Figure A with interpreted depositional environment added. 69

Error! Not a valid link.

OC1-B-St OC1-M-St OC5-MK-St OC3-RLB1 OC5-RLB1 Mineral Count % Count % Count % Count % Count % Monocrystalline Quartz 292 48.7 304 50.7 214 35.7 138 23.0 69 11.5 Polycrystalline Quartz 158 26.3 115 19.2 110 18.3 53 8.8 45 7.5 Pore 0 0.0 2 0.3 4 0.7 1 0.2 0 0.0 Na-Ca Feldspar 32 5.3 28 4.7 27 4.5 8 1.3 6 1.0 Biotite 1 0.2 49 8.2 11 1.8 13 2.2 1 0.2 Muscovite 24 4.0 37 6.2 13 2.2 27 4.5 12 2.0 Limonite 0 0.0 0 0.0 13 2.2 2 0.3 1 0.2 K-Feldspar (Orth) 0 0.0 0 0.0 0 0.0 39 6.5 27 4.5 Garnet 9 1.5 1 0.2 1 0.2 0 0.0 0 0.0 Opaque 5 0.8 4 0.7 2 0.3 4 0.7 5 0.8 Quartz Matrix 78 13.0 60 10.0 29 4.8 37 6.2 4 0.7 Calcite Matrix 0 0.0 0 0.0 176 29.3 264 44.0 428 71.3 Zircon 0 0.0 0 0.0 0 0.0 0 0.0 2 0.3 Chert 1 0.2 0 0.0 0 0.0 0 0.0 0 0.0 Unknown 0 0.0 0 0.0 0 0.0 14 2.3 0 0.0

Total 600 100 600 100 600 100 600 100 600 100

Q 450 93.2 419 93.7 324 92.3 191 80.3 114 77.6 F 32 6.6 28 6.3 27 7.7 47 19.7 33 22.4 L 1 0.2 0 0.0 0 0.0 0 0.0 0 0.0 Total 483 447 351 238 147

Qm 292 60.5 304 68.0 214 61.0 138 58.0 69 46.9 F 32 6.6 28 6.3 27 7.7 47 19.7 33 22.4 Lt 159 32.9 115 25.7 110 31.3 53 22.3 45 30.6 Total 483 447 351 238 147

Table 4.2: Results from point counting of 5 Facies 4a sandstone samples.

70

Figure 4.27: Results of point counting plotted on ternary diagrams from Dickenson et al., 1983. Colored stars on the ternary diagram and photomicrographs coincide with their location on the generalized section on the right. Thin sections were made for each of the of the major sandstone units of the study interval in order to view the framework mineralogy changes up-section. The “Black Star”, erosional sandstone was identified in the field as a lithic arenite and the basal (Green Star) was a quartz arenite however point counting results showed that they were both quartz arenite with the basal sandstone having more polycrystalline quartz. Results show that two out of the three calcite cemented sandstones were lithic arenites and had “Mixed” provenance signatures on the QmFLt diagram which is uncommon for sandstones in the Appalachian Basin. Most lie in the Quartzose Recycled field. It is possible that the data is skewed because in OC3-RLB1 and OC5-RLB1 44% and 71% of the counts were calcite cement. Although OC1-MK- St had 29% and was still a quartz arenite and plotted in the Quartzose Recycled cell. 71

Figure 4.28: Rose diagram showing the results from the paleocurrent analysis of a Facies 4a sandstone. Braided streams give bidirectional patterns however the exposures in this outcrop exhibited two sided troughs, where the entire trough was exposed, measurements along the trough hinge is likely the cause for the dissecting vector and is there for the true paleocurrent direction in the sampled location.

72

Figure 4.29: Lateral accretion surfaces observed in the erosional Facies 4a sandstone within OC-1. These surfaces are parallel to the basal erosional contact.

1m

Figure 4.30: Planar tabular crossbedding observed in OC-5 with sharp basal and upper contacts.

74

4.2.5 Facies 5: Coal

Description: The three organic beds within the study interval, the Lower Kittanning (LK),

Middle Kittanning (MK), and the Upper Freeport (UF) coals are only half of those that ere mined within the Allegheny Group in Ohio (Sturgeon, 1958). The Brookville (No. 4), and

Clarion (No. 4a) are missing at the base and the Upper Kittanning (No. 6a) and Lower

Freeport are absent near the top. Each of the coals present occurs above a blocky mudstone (Facies 2c) and below a black shale (Facies 1) or sandstone (Facies 4a). The basal contacts with the blocky mudstones are abrupt whereas the upper contacts vary from gradational into black shale (Facies 1) to erosional into sandstone.

The LK coal is present in four of the six outcrops and has an average thickness of 0.6 meters (ranging from 0.52 – 0.64 m) throughout the study area. The Middle

Kittanning (MK) coal is present in three of the sections and has an average thickness of

1.5 m. In OC-5 the MK coal is absent due to erosion by the overlying sandstone. The coals in these sections have columnar structures, a yellow clay with a sulfur odor, and charcoaled and carbonized plant fragments (Figure 4.31). The Upper Freeport (UF) coal is present only in OC-4.

Interpretation: The coals are considered to have formed in tropical climates within raised mires (Cecil et al., 2003; Greb et al., 2003). The regional extent of the coals combined with the location above paleosols of Facies 2c and below laminated mudstones of

Facies 2b indicate they were formed during transgressions as eustatic sea level change raised regional base level (Heckel, 1994). Mack et al. (1993) considered any coal accumulation a Histosol.

75

30cm

Figure 4.31: Middle Kittanning coal bed exhibiting a columnar structure with a sharp basal contact overlying a blocky mudstone.

76

4.3: Facies Associations

Seven facies associations were identified in this study, they were defined based on interpreted differences in depositional environment of facies and groups of facies. Facies associations are grouped into terrestrial or marine and further subdivided into specific depositional environments (Table 4.3). Defining depositional environments via facies associations was an integral step in the creation of the sequence stratigraphic framework and understanding of depositional controls on sedimentation and sediment preservation.

77

Facies Lithofacies Environment Association a 2a Black, fossiliferous mudstone Prodelta Marine FA-1 1 Black shale b Prodelta 3 Laminated siltstone 1b Gray laminated mudstone Lake Subaqueous 3 Laminated siltstone FA-2 Delta 4d Horizontally bedded sandstone 2c Blocky mudstone FA-3 Delta Plain Interfluve 5 Coal 4a Trough crossbedded sandstone Continental Single-Story 4c Rippled sandstone a Sandstone 4d Horizontally bedded sandstone 4a Trough crossbedded sandstone Fluvial FA-4 Multistory 4b Planar tabular crossbedded sandstone Channel b Sandstone 4c Rippled sandstone 3 Laminated siltstone Abandoned c 2b Gray, laminated mudstone Channel Table 4.3: Facies associations were created from grouped lithofacies then interpreted in terms of depositional environment.

78

4.3.1 FA-1: Pro-Delta Mudstones and Shales

The dark, organic mudstones and shales of FA-1 represent the lowest energy siliciclastic environments in the study area. Prodelta sediments are a combination of deposition from suspension and hyperpycnal flows (Bates, 1953). The relative contribution of each component depends on the salinity contrast between the river and basin water, the concentration of suspended load, and distance from the river mouth.

Prodelta sediments occur in both marine and lacustrine environments and can be differentiated in modern sediments by faunal content. In the study area both fossil content and grain size were used to subdivide this facies association into marine and lacustrine environments.

4.3.1.1 FA-1a Marine Prodelta

Description: FA-1a consists fossiliferous, black mudstone facies (Facies 2a) at the base of Allegheny Group in section OC-1 and OC-6.

Interpretation: The abundance of organic matter and fine grain size of this facies association suggests that these sediments were in a prodelta setting with rapid deposition that included a high volume of organic matter. The sharp basal contact is interpreted to be a wave ravinement surface over terrestrial landscape based on the occurrence of the mudstone-filled burrows of the Glossifungites ichnofacies penetrating the underlying paleosol (FA-3). The abundance of fossils in the basal 10 cm indicates the presence of a condensed section. No bioturbation was identified in this lower section, however the lack of distinct lamination at the base may be a result of infaunal reworking. The abundant siderite concretions with marine fossils on the surfaces formed during early diagenesis after deposition in a normal marine setting. The abundance of 79 siderite concretions and scarcity of pyrite suggests rapid sedimentation in shallow water depths (< 10 m) (Gautier, 1982). A shallow water depth is consistent with the mudstone as the landward equivalent of the Putnam Hill Limestone (Merrill, 1950), but inconsistent with the fauna present.

4.3.1.2 FA-1b Lacustrine Prodelta

Description: FA-1b consists of black laminated shales of Facies 1 that are found overlying coals in sections OC-2, 4 and 5. The black shales are finer grained than the mudstones of FA-1a sediments and lack body fossils and ichnofauna. Fossils have been reported from stratigraphically equivalent deposits above the Middle Kittanning and

Lower Freeport coals, termed the Washingtonville and Dorr Run shales respectively.

The basal contacts are gradational with the underlying coals and upper contacts also gradational into laminated gray mudstones and siltstones of FA-2.

Interpretation: The lack of body or trace fossils, thin laminations, high organic content is indicative of stressed environmental conditions in FA-1b, at least locally. The gradational contact at the base of the shales is interpreted as either the result of progressive drowning of a peat mire (FA-3) (e.g., Wright, 1979) or the erosion of organics from the top of the mire by wave action during a more rapid water level rise. The lack of wave ripples suggests that the progressive drowning scenario is more likely. The gradational upper contact formed when hyperpycnal flows began filling the lake basin.

4.3.2 Facies Association 2: Subaqueous Delta

Description: FA-2 is composed of interbedded mudstones (Facies 2b) and siltstones

(Facies 3) and thin, horizontally bedded sandstones (Facies 4d). FA-2 lies above the 80 unfossiliferous mudstones of FA-1 and is capped by blocky mudstones of FA-3. This facies association occurs below the Lower Kittanning coal in sections OC-2 and OC-3, between the Lower and Middle Kittanning Coals in section OC-4 and OC-5, and above the Lower Freeport Sandstone in section OC-4. The there is no single vertical facies pattern, but in all cases the relative water depth first increases and then decreases. In the study area siderite nodules are common in FA-2, but the only fossil material present was carbonized plant matter.

Interpretation: Deltaic sediments deposited on the prodelta can be subdivided into delta front, interdistributary bays and distributary mouth bars (DMB) (e.g. Tye and Coleman,

1989a,b). Delta Front sediments are often depicted as coarsening upward progradational systems (e.g., Olariu and Bhattacharya, 2006) that contain less fauna than the prodelta setting. Interdistributary bays are complex depositional systems, which can resemble any element of a deltaic system, e.g., prodelta, delta front, distributary channel, and levee (Tye and Coleman, 1989a).

Cores from modern lacustrine deltas on the Mississippi Delta plain show that interdistributary regions accumulate fine sediments that have no distinctive vertical pattern. The extension of distributary channels during basin fill in lacustrine and marine deltas leads to the formation of DMB sand deposits (Tye and Coleman, 1989a), which in this study area are siltstone to very fine-grained sandstone. DMB sandstones are depicted by different authors as thin lenses (Elliot, 1986) and as more sheet-like (Olariu and Bhattacharya, 2006) with variable thickness. The DMB sandstone and siltstone beds shown in the cores of the Atchafalaya Delta (Tye and Coleman 1989a) are only a few decimeters in thickness but up to several km in lateral extent. 81

A deltaic interpretation for sediments of the Allegheny Group is not new. Most of the literature of the 1970s and 1980s (Ferm 1970, Donaldson, 1985) concluded that the

Allegheny Group represent a fluvially dominated deltaic system analogous to the modern

Mississippi. A similar expression of FA-2 is seen in the coarsening upwards sequence of Appalachian basin cyclothems described by Chesnut, 1994 (Figure 4.32). The

Breathitt succession in which he based this interpretation is equivalent to the Allegheny

Group deposit of this study. Chesnut termed these strata coal-clastic cycles and interpreted the coarsening-upwards sections as marine or brackish water environments.

82

Figure 4.32: Generalized Appalachian cyclothem model modified from Chesnut, 1994. In this interpretation the clastic cycle begins with a marine or brackish shale above the coal bed and is followed by a coarsening upwards profile made of interbedded siltstone, mudstones and sandstones. These coarsening upwards deposits are the FA 2 sediments in this study. The coarsening upwards sediments are erosionally truncated by a new, fining upwards sequence which is equivelent to the fluvial channel fill in this study’s depositiopnal framework, capped by a blocky mudstone and coal bed.

83

4.3.3 Facies Association 3: Delta Plain Interfluve

Description: FA-3 is chiefly composed of the Facies 2c (Blocky Mudstones) and coal

(Facies 5). Coal is present above paleosols in three of the six interfluve horizons recognized in the study interval. This facies association has gradational lower contacts with FA-2 or FA-4 and sharp upper contacts with FA-1 prodelta sediments or FA-2 mudstones. FA-3 is found in all measured sections.

Interpretation: An interfluve is defined as the region between fluvial channels (Horton

1945). Interfluve regions may, therefore, be local or regional in extent and have erosional and aggradational sediment patterns (Gibling, 2006). The interfluves formed when falling base level forced the distributary channels of the delta to incise (Chapter 5).

Subaerial exposure led to the onset of pedogenic processes by meteoric water percolation, bioturbation, and wetting-drying. A rise of base level led to gleying of most of the soil profiles. The correlation of the coal seams within and between basins indicates that the interfluves were laterally extensive as well.

The vertical stacking pattern of the paleosols and their overlying facies provides an indication of the rate of burial. Paleosols capped by coals, which in this case are laterally extensive, indicate a rise in base level that is regional in extent and likely the result of a glacial-eustatic marine transgression (Heckel, 2008). The paleosol horizons not capped by coals are more variable. The paleosol underlying the marine transgression at the base of the section is reported to be capped by a coal elsewhere

(Sturgeon, 1958). However, in the sections of this study, the paleosol lies disconformably under a marine mudstone. One explanation for the differences is local topographic variations. It is unlikely that no incision of the interfluve occurred during exposure. During base level rise coals formed in lower areas but not over the higher 84 elevations. The absence of coal above blocky mudstones higher in the section within the study can be explained by local topographic variations and equally local erosion by distributary channels incision when base level fell.

85

4.4 Fluvial Channel Deposits

Fluvial channels occur as both single (FA-4a) and multistory (FA-4b) lithosomes with sharp to erosional bases and variable amounts of incision. Two examples of fine- grained channel deposits (FA-4c) occur near the top of section OC-4.

4.4.1 FA-4a Single Story Sandstone

Description: Three single story sandstones of FA-4a are present in section OC-1, OC-2,

OC-4 and OC-6. The lower contact of the sandstone at the base of OC-1 is not exposed. The sandstones are composed of Facies 4a overlain by Facies 4b. The sandstones within section OC-2 and OC-4 have a sharp to erosional base and are composed of trough cross-bedded sandstone (Facies 4a), horizontally laminated sandstone (Facies 4d) and rippled very fine-grained sandstone (Facies 4c) in an overall fining upward succession. Both examples overlie the Middle Kittanning Coal. The example in section OC-6 has an erosional base and fines upward only slightly, although much of the upper portion of this sandstone was inaccessible. The only structures visible were trough crossbeds of Facies 4a.

Interpretation: FA 4a was deposited by fluvial channels. The style of fluvial system is determined partially by the sandstone lithosomes and partly by the adjacent facies (Miall,

2013). In the study area, the only component that can be used is the channel sandstone. Based on the relative abundance of planar tabular and trough cross beds the basal sandstone of OC-1 is interpreted as a braided fluvial system (e.g., Miall 1977).

The abundance of Facies 4a in the remainder of the single-story sandstones suggests deposition by either anastomosed or meandering fluvial channels.

86

4.4.2 FA-4b Multistory Sandstone

Description: One multistory sandstone occurs near the base of the Allegheny Group and is present at the top of sections OC-1 and OC-5 and the base of sections OC-2, OC-3 and OC-6. The sandstone varies from 7.6 – 10.5 m in thickness. The base of the sandstone in section OC-1 has an intraclast horizon of sideritized shale clasts and dark gray mudstone of Facies 2a along an erosional surface with up to 5 m of relief. In the basal portion of the FA-4b sandstone in OC-5 there is a 1.5 m wide, 30 cm thick mat composed of coal and carbonaceous, laminated mudstones within a fluvial sandstone unit. This mat is thought to have been excavated from the underlying FA-1 and FA-3 and transported by a high energy fluvial system (Figure 4.33). The base of the sandstone in section OC-1 displays lateral accretion surfaces that drape the erosion surface.

Individual sandstones are delineated by reactivation surfaces which are either may have layers of mudstone or intraclasts on the surfaces. Internal structures within the sandstones are dominated by trough crossbedding with subordinate planar tabular crossbedding, horizontal lamination and ripple cross-lamination at the top. The trough crossbeds in sections OC-3 show flow toward 330°. The width/thickness of the upper story perpendicular to paleoflow is a minimum of 122 m.

Interpretation: The origin of the multistory sandstone is open to several interpretations.

If the sandstone is essentially syndepositional then it could represent a distributary mouth bar deposit of a major distributary (Tye, and Coleman, 1989a; Olariu, and

Bhattacharya, 2006). Alternatively, the reactivation surfaces could mark the superposition of channels belonging to different incision events. At the study area there are no data available to distinguish between the two possibilities. However, the combination of the presence of mineable coals within the interval spanned by the 87 sandstone in counties to the south (Stout, 1923) and the lack of a significant depositional hiatus indicated by the paleosols present suggests the multiple incision scenario is more likely. In either case the infill of the channels was by withe ran anastomosed or meandering fluvial system.

4.4.3 FA-4c Abandoned Channel

Description: Two fining upward successions composed of siltstone (Facies 3) and laminated gray mudstones (Facies 2b) are present near the top of section OC-4. The lower example rests on sandstones of FA-4a and the upper on a paleosol of FA-3. Both examples have sharp basal contacts and gradational upper contacts into FA-3.

Interpretation: The fining upward grain size pattern indicates decreasing energy which is consistent with a channel abandonment. Such abandonments occur on modern delta plains (Tye and Coleman, 1989 a,b). The problem lies in the absence of a sandstone lag at the base which is commonly present, even if thin (Miall, 2013). An alternative is an interdistributary trough (Tye and Coleman, 1989b), which only receives fine-grained sediment from flooding events.

4.4.4 Discussion

Fluvial channel sandstones are commonly assigned to one of three possible styles; braided, meandering, or anastomosing (Schumm, 1981). Braided fluvial systems consist of bedload channels floored by sinuous-crested bedforms that form trough crossbeds in the deepest portions of channels adjacent to positive topographic features termed bars. The migration of bars laterally or downstream forms planar tabular crossbeds that overlie the trough crossbeds (Miall, 1977). A basal lag may or may not 88 be present and finer grained material is present only as rare caps on the bars. As a result, grain size within channel deposits decreases only slightly up-section.

The structures with meandering fluvial channels are a product of unidirectional flow mixed load rivers and the lateral migration of point bars (Schumm, 1981). The channels generally have a basal lag of intraclasts formed as the cut-bank erodes the associated floodplain mudstones. The grain size and structures within the sands vary upward from sinuous crested bedforms to rippled sand higher in the section as both velocity and grain size decrease up the point bar (Walker, 1976). The migration of the point bar forms the characteristic lateral accretion deposits (Walker and Cant, 1984).

Gravitational settling of grains in these systems results in a pronounced fining upward of grain size.

Anastomosed fluvial deposits are formed by suspended-load rivers (Schumm,

1981). In this fluvial style the channel pattern is also sinuous but there is almost no lateral migration due to the low energy of the fluid flow and the inter-channel lacustrine and marsh mudstone deposits. Sinuous crested bedforms occur in the channels

(Nadon, 1994). The characteristic geometry of the fluvial channels is a lenticular shape with levee deposits forming 'wings' (Nadon, 1994). The floodplain sediments of anastomosed systems are volumetrically larger than the channel sands. The floodplain sediments are interbedded lacustrine and marsh mudstones and crevasse splay sandstone sheets (crevasse splays).

The erosive base and fining upward grain size indicate that the single story sandstones of FA-4a represent the deposit of single fluvial channel events. The coarse grain-size, which is not seen within the underlying and overlying deposits, indicates that the bedload carried by the system was not transported into the adjacent depositional 89 flood basin. The capture of the coarse-grained material may be a result of rapid aggradation of the fluvial system as a whole or the infill of an incised channel.

The sandstones with reactivation surface of FA-4b represent amalgamated fluvial sandstones. The superposition of multiple channels in only one part of the sections is interpreted to be caused by a lower than average subsidence rate at the time of deposition. The fine-grained channel deposits of FA-4c are interpreted to be the result of decreasing competence as a result of channel abandonment.

4.5 Depositional Model – Fluvially Dominated Delta

The marine shales within the Allegheny Group (FA-1) were deposited as a prodelta by a super-estuarine marine system characterized by large-scale, estuarine- type circulation in combination with lateral influx of oxygen-deficient intermediate waters of the Late Pennsylvanian Midcontinent Sea (Figure 4.34); Heckel, 1977, Algeo and

Heckel, 2008). When compared to the midcontinent cyclothems, the decrease in the number of biostatigraphically constrained marine units and well-developed paleosols suggests that the strata in the study area and in the Appalachian basin were deposited higher on the shelf in shallower water (Figure 2.4) (Heckel, 1995).

The bulk of the section is interpreted as lacustrine delta. The deltas of modern

Atchafalaya River (Tye and Coleman, 1989a,b) are considered to be a good modern analogue for the present study section because of the limited total accommodation available in both locations (Fig. 4.35). Facies patterns are similar in bed thickness and grain size. The primary difference is that the distributary channels also incised the delta during multiple sea level lowstands towards the north-west (Chapter 5, Figure 4.36).

90

Figure 4.33: Raft composed of coal and carbonaceous laminated mudstone in section OC- 5 interpreted to be excavated and transported by high energy fluvial systems. 91

Figure 4.34: Depiction of vertical circulation within a west facing, tropical, epicontinental sea, such as the LPMS. Modified from Heckel 1977.

92

200 m

Figure 4.35: Google Earth image of the Atchafalaya River drainage system which acts as a modern analogue for sediment deposition in the Allegheny Group within the study area.

93

Figure 4.36: Paleocurrent azimuth (330°) represented by the red arrow superimposed on the Middle Pennsylvanian paleogeographical maps from Donaldson et al.(1985). Study area is denoted by the red circle.

94

CHAPTER 5

SEQUENCE STRATIGRAPHIC FRAMEWORK

5.1 Introduction

The development of sequence stratigraphy was an important advancement to the prediction of environments of depositions and their spatial distribution (Van Wagoner et al., 1988). The interpretation of seismic images from passive margins in terms of packages of strata bounded by unconformities allowed thick sections of the stratigraphic column to be organized in a chronostratigraphic framework based solely on stratal geometry (Payton, 1977). Exxon then extended this approach to much smaller scales with the development of sequence stratigraphy (e.g., Van Wagoner et al., 1990). After a brief review of sequence stratigraphic terminology, the study area is interpreted in terms of changes in relative sea level and then in terms of controls on sedimentation.

5.2 Sequence Stratigraphy Terminology

Unconformity bounded packages of sediment, which were first termed sequences by Sloss (1963), are the fundamental units of sequence stratigraphy.

Sequences are defined as relatively conformable succession of genetically related strata, bounded at the top and base by unconformities and their correlative conformities

(Mitchum, 1977, Vail et al.,1977). The original concepts of sequence stratigraphy are based on the balance of accommodation and sediment supply. Eustatic sea level change and tectonic subsidence or uplift are the main controls on accommodation, i.e., the space available for sediment deposition between sea level and the sea floor (Jervey,

1988). Total accommodation refers to the amount of space available and determines the thickness of deposits. The rate of formation of accommodation describes the rate at 95 which relative sea level rises or falls, mostly due to tectonic subsidence/uplift and eustasy. The rate of formation of accommodation controls the distribution of environments. The term relative sea level is commonly used to refer to the apparent rise or fall of sea level based on sediment patterns, such as fining upward or coarsening upward, that can be generically interpreted as transgression or regression without knowing the cause for the change.

The concepts of sequence stratigraphy were first and most commonly applied to passive margin deposits. Sequences are separated from one another by sequence boundaries (SB), which forms in response to relative sea level fall (Van Wagoner et al.,

1990). The fundamental unit within a sequence is the parasequence, which is defined as the smallest progradational unit bounded by marine flooding surfaces (Figure 5.1).

Parasequence sets form a distinctive vertical stacking pattern and that can be progradational, retrogradational, or aggradational (Figure 5.1) (Van Wagoner et al.,

1990).

A progradational parasequence set forms a thickening and coarsening upward package deposited when the rate of sediment input exceeds the rate of formation of accommodation causing a relative sea level fall. It should be noted that absolute sea level is still rising otherwise the negative total accommodation would result in a depositional hiatus or erosion surface rather than a deposit. If the rate of formation of accommodation is greater than the rate of sedimentation, then total accommodation will increase and the result is a landward shift of facies over time forming a thining and fining upward succession termed a retrogradational parasequence set. If sediment influx equals the rate of formation of accommodation then there will be no change in relative sea level over time and an aggradational parasequence set forms. 96

Sediments deposited contemporaneously, such as lagoon, beach, shelf, deep sea, form depositional systems (Fisher and McGowen 1967, in Van Wagoner et al.,

1990). The response of depositional systems to changes in the rate of accommodation forms systems tracts, which is the term used to describe sediments deposited under the same conditions of relative sea level change. Systems tracts are organized based on types of bounding surfaces, parasequence set distribution, and position within a sequence and can also be defined by geometry and facies associations.

There are three basic systems tracts, Lowstand, Transgressive, and Highstand.

The lowstand systems tracts (LST) refers to sediments deposited after a time of relative sea level fall when relative sea level is low and slowly rising. The sediments in the LST form a progradational parasequence set. On a passive margin the LST is deposited below the shelf break and has no chronostratigraphic equivalent strata on the shelf (Van

Wagoner et al., 1990). In a foreland basin setting the location of the LST depends on the magnitude of absolute sea level fall. There may be a thin LST (e.g., Van Wagoner et al., 1990, fig 20a) or none at all.

The transgressive systems tract (TST) describes retrogradational sediments deposited during relative sea level rise after the LST. The TST directly overlays a transgressive surface (TS) and is capped by a maximum flooding surface (MFS). The

MFS represents the highest point of relative sea level rise. Flooding surfaces are common tools of stratigraphers as they are easily recognizable, widely distributed and can represent significate events. The upward fining of sediments in the retrogradational systems tract reflects increasing distance to sediment source. The decrease in sedimentation rate results a thin deposit that may represent a long periods of time termed a condensed section (CS) (Loutit et al., 1988). The highest diversity and fauna 97 are found at these condensed sections which make them important for biostratigraphy and fauna studies.

The highstand system tract (HST) is bounded on the base by the MFS and on the top by the upper sequence boundary. The HST is composed of a progradational parasequence set deposits as the rate of formation of accommodation is slowing.

Assuming sedimentation rate stays constant, the result is a regression of the shoreline.

Additional systems tracts are defined in areas where the basin style and geometry produce variations in accumulation rate and thickness. In regions of high total accommodation, such as passive margins, a Shelf Margin Systems Tract forms as absolute sea level falls to the shelf edge (Van Wagoner et al., 1990). In areas of low total accommodation, such as foreland basins, a falling stage systems tract (FSST) can be recognized in which isolated shoreface sand bodies are formed in response to a forced regression (Plint and Nummedal, 2000). In a distal foreland basin setting during a period of high magnitude, high frequency eustatic sea level fluctuations the expected systems tracts include the TST, HST with the LST, FSST being expressed as sequence bounding paleosols or surfaces of non-deposition (Martino, 2016).

5.3 Allegheny Group Sequence Stratigraphic Model

In this study a sequence stratigraphic framework has been constructed and applied based on the vertical distribution of facies associations (Figure 5.2). A total of 10 unconformities defining nine sequences were found interpreted to be present within the

Allegheny Group (Figure 5.3). These unconformities are interpreted to be at the base of fluvial channels (FA-4) or the tops of paleosol horizons (FA-3).

There are three systems tracts recognized in the field area, an early transgressive systems tract (eTST), a transgressive systems tract and a highstand 98 system tract. The eTST is composed of fluvial channel deposits (FA-4), are 4 - 15 m thick, and span the entire field area. However, eTST thickness can vary up to 5 m due to the erosive nature of the basal contact. TSTs overly the fluvial channel deposits of the eTSTs and also the blocky mudstones of FA-3. TSTs are made up of the coal beds of

FA-3, prodelta sediments of FA-1and the siltstones and gray, laminated mudstones of

FA-4c. TSTs within the study area are thin compared to the eTST and HST strats. TST thickness ranges from 0.10 – 3m and spans the study area. However, in multiple areas

TST coals and shales are erosionally truncated by eTST sandstone bodies (Chapter 4,

Figure X). HST deposits occur above the interpreted flooding surfaces or at the top of the TST. They include the upper portion of the FA-1 prodelta beds, and interbedded mudstones and siltstones of FA-3. HST thickness in the study area ranges from 0.75 – 6 m and lateral extent spans the field area.

5.4 Discussion

The key aspects to the proposed model is the combination of high magnitude, high frequency glacial-eustatic sea level changes, a region of low and variable total accommodation, and a deltaic depositional system. A drop in base level during eustatic sea level fall resulted in incision of the fluvial distributary channels and the exposure of the interfluves where the paleosols of FA-3 were formed. There are no equivalents to the LST deposits in the field area. The LST deltas for this time period occur within the

Strawn interval of the Permian Basin of (Wright, 1979 and Heckel, 2008) and therefore the onset of eustatic sea level rise was not recorded in the deposits of the study area (Figure 2.4).

The limited degree of incision of the distributary channels is a result of initial downcutting as the sea retreated that was then halted by intermediate base levels. The 99 magnitude of sea level fall during the Middle Pennsylvanian was on the order of 40 m

(Rygel et al., 2008), however, the response of the Mississippi River to a drop in sea level of twice the magnitude was to entrench the main channel approximately 600 km upstream over a period of approximately 100 kyr (Shen et al., 2012). The duration

Pennsylvanian glacial cycles have been estimated to vary from 100 kyr (Maynard and

Leeder, 1992) to 400 kyr (e.g. Heckel, 2008). Even with a 400 kyr cyclicity it is unlikely that the smaller Middle Pennsylvanian sea level fluctuation could have resulted in the

Lowstand incision knickpoint migrating from the Permian basin to the Appalachian basin.

The initial response to the rise in base level in the study area is represented by the infilling of distributary channels (FA-4) during early transgressive systems tract

(eTST) (Figure 5.2). During higher magnitude transgressive events the continued rise in base level (TST) led to the formation of a flooded delta plain and the development of peat mires on the interfluves once the channels were filled. During the largest transgressions the peats were buried by marine pro-delta sediment (FA-1) or eroded by wave action. The limited fauna within the marine shales of the Allegheny Group (FA-1a) may be an effect of the estuarine-type circulation with lateral influx of oxygen-deficient intermediate waters postulated by Algeo and Heckel (2008) for the Appalachian basin.

Less extensive transgressions resulted in the peat mires being covered by thin lacustrine pro-deltaic deposits.

The HST sediments are composed of a complex assemblage of delta front and interdistributary bay sediments of FA-2 that prograded into the area as the rate of sea level rise decreased. The limited total accommodation was responsible for the formation of thin, rapidly prograding delta complexes. Avulsions, which are common on Highstand

Deltas (e.g., Tye and Coleman 1989a,b) would have been stopped by channel entrenchment as soon as eustatic sea level fell. 100

The regional differences between the stratigraphy of the Allegheny Group in this study area and those along strike in the Appalachian basin is interpreted to be a result of a combination of channel avulsion and lateral variation in the amount of accommodation formed during and between transgressive events. A comparison of the composite stratigraphic column to the Mid-Continent section compiled from Boardman et al. (2004) and Heckel (2013) (Figure 2.3) shows that even some of the significant marine incursions are absent in the study interval suggesting that subsidence rates were variable.

101

Figure 5.1: Stacking pattern of progradational, retrogradational and aggradational parasequence sets on a passive margin, displayed in cross section and schematic log responses. Figure from Van Wagoner et al., 1990.

102

Figure 5.2: Generalized, representative measured section and sequence stratigraphic interpretation with relative 4th order sea level curve. Light green lines represent flooding surfaces or period of maximum relative sea level rise. Red lines are unconformities mostly caused by glacio-eustatically driven falls in sea level. Facies associations are labeled on the stratigraphic section and relative sea level curve.

103

Figure 5.3: Composite measured section with sequence stratigraphic interpretations including 4th order relative sea level curve, systems tract and depositional environment based on facies associations.

104

CHAPTER 6

CONCLUSIONS

The results of this study are

1) The 4th-order eustatic sea level changes played a major role in sequence formation but that the sediment response in the back-bulge region was sensitive to spatially and temporally variable subsidence rates.

2) Both XRD and XRF data show that the anomalously thick Lower Kittanning clay bed was formed by amalgamation of two paleosol profiles.

3) The petrography of the sandstones and the paleocurrent data are consistent with deposition by river systems flowing to the northwest from the Appalachian Mountains.

The difference in provenance regime may be evidence for the sandstones being deposited in different cycles. The Brookville Sandstone has a similar framework mineralogy to the sandstones of the Allegheny Group however there is a difference in cement. Based on hand sample observation, the basal sandstone has a much higher clay content in the matrix than the sandstones in the rest of the section in the field area.

Also, the Brookville sandstone lays below the Putnam Hill shale, therefore, it is considered to be the uppermost lithosome of the Pottsville Group rather than the basal sandstone of the Allegheny Group.

4) The overall depositional environment was one of a series of thin, mainly lacustrine deltas deposited as Highstand Systems Tracts. The prograding deltas were similar in 105 size and thickness to those forming in the modern Atchafalaya Basin (Tye and Coleman,

1989a,b). The grain size and thickness of the channel sandstones indicates that they incised through older deltaic deposits during glacial lowstands and were back-filled during the early Transgressive System Tract. Most of the paleosols formed on interfluves isolated from sediment influx during lowstand. The exception is the composite paleosol beneath the Lower Kittanning coal, which could indicate the presence of an additional sequence boundary or burial of the older soil profile by sedimentation on the interfluve (e.g., Gibling et al., 2005).

5) The paleosols in the section differ from base to top in the study interval. All the paleosols are dominated by kaolinite. The structures and clay composition of the basal paleosol indicates formation as possible eutric Argillisol (Mack et al. 2003; Sheldon and

Tabor, 2009) whereas the paleosols below the Lower Kittanning and Upper Freeport coals are paleoVertisols (Mack et al., 1993). Classification of the basal paleosol is less definitive because of erosion of the upper portion by wave action during transgression.

6) The weathering indices calculated from XRF bulk geochemical analysis of the paleosol profiles show the CIA-K have a minimum of 96 and maximum of 99, these values are at the high end of the CIA-K spectrum and indicate a high degree of weathering caused by the kaolinization of feldspars and aluminosilicates, either from meteoric water percolation or the effects of a sustained high water table due to high base level and poorly drained conditions. Average paleoprecipitation values range from

1347-1374 mm/year using the CIA-K weathering index and 1590 - 1714 mm/year using

CALMAG. The estimated MAP values indicate a small increase from the base to the

Lower Kittanning interval and then a decrease to the Upper Freeport paleosol. This 106 slight, overall trend is consistent with the climate model of a shift to a more seasonal climate due to a rain shadow affect from the growing Appalachian mountains.

107

REFERENCES

Algeo, T.J., and Heckel, P.H., 2008, The Late Pennsylvanian Midcontinent Sea of North America: a review, Palaeogeography, Palaeoclimatology, Palaeoecology, v. 268 (3), p. 205-221. Allen, J.R.L., 1984, Parallel lamination developed from upper-stage plane beds: a model based on the larger coherent structures of the turbulent boundary layer, Sedimentary Geology, v. 39, p. 227-242. Baird, G.C., Sroka, S.D., Shabica, C.W. and Kuecher, G.J., 1986, Taphonomy of Middle Pennsylvanian Mazon Creek area fossil localities, Northeast Illinois: Significance of exceptional fossil preservation in syngenetic concretions, Palaios, v. 1 p. 271- 285. Bates C. C., 1953, Rational theory of delta formation, Bulletin of the American Association of Petroleum Geologists, v. 37 (9), p. 2119-2162. Berner, R. A., 1971, Worldwide sulfur pollution of rivers, Journal Geophysical Research, v. 76, p. 6597-6600. Blakey, R. C., 2007, Carboniferous-Permian paleogeography of the assembly of Pangea, Fifteenth International Congress on Carboniferous and Permian Stratigraphy, in, T.E. Wong, (Ed), Royal Netherlands Academy of Arts and Sciences, Amsterdam, Utrecht, the Netherlands, p. 443–465. Blakey, 2011, Paleogeography, Website accessed 4 May 2016, http://www2.nau.edu/rcb7/nam.html. Boardman, D.R., Heckel, P.H., and Marshall, T.R., 2004, Preliminary report on lower Desmoinesian (mid-Moscovian) from lower and middle Cherokee Group of southern Midcontinent North America, Newsletter on Carboniferous Stratigraphy, v. 22, p. 41-47. Busch, R. M., and Rollins, H. B., 1984, Correlation of Carboniferous strata using a hierarchy of transgressive-regressive units, Geology, v. 12, p. 471-474. Catuneanu O., Zecchin M., 2013, High-resolution sequence stratigraphy of clastic shelves II: Controls on sequence development, Marine and Petroleum Geology, v. 39, p. 26-31. Cecil, C.B., 1985, Paleoclimate controls on Late Paleozoic sedimentation and peat formation in the central Appalachian basin (U.S.A.), International Journal of Coal Geology, v. 5, p. 195–230. Cecil, C.B., 1990, Paleoclimate controls on stratigraphic repetition of chemical and siliciclastic rocks, Geology, v. 18, p. 533–563. Cecil, C.B., 2003, The concept of autocyclic and allocyclic controls on sedimentation and stratigraphy, emphasizing the climate variable, Society for Sedimentary Geology Special Publication 77 p. 13–20.

108

Cecil, C.B., 2013. An overview and interpretation of autocyclic and allocyclic processes and the accumulation of strata during the Pennsylvanian–Permian transition in the central Appalachian Basin, USA, International Journal of Coal Geology, v. 119, p. 21–31. Cecil C. B., and Dulong, F. T., 2003, Precipitation models for sediment supply in warm climates, in, C.B. Cecil and N.T. Edgar, (Eds), Climate controls on stratigraphy, Society for Sedimentary Geology Special Publication 77, p. 21–28. Cecil, C.B., and Edgar, N.T., 2003, (Eds), Climate controls on stratigraphy, Society for Sedimentary Geology, Special Publications 77, p. 275. Cecil, C. B., Stanton, R. W., Neuzil, S. G., Dulong, F. T., Ruppert, L. F., and Pierce, B. F., 1985, Paleoclimate controls on Late Paleozoic sedimentation and peat formation in the central Appalachian basin, U.S.A., International Journal of Coal Geology, v. 5, p. 195–230. Cecil, C.B., Dulong, F.T., West, R.R., Stamm, R., Wardlaw, B., and Edgar, N.T., 2003, Climate controls on the stratigraphy of a Middle Pennsylvanian cyclothem in North America, Climate Controls on Stratigraphy, Society of Sedimentary Geology, Special Publication No. 77, p. 151–180. Cecil, C.B., Brezinski, D.K., and Dulong, F.T., 2004, The Paleozoic record of changes in global climate and sea level: Central Appalachian Basin, in, S. Southworth and W. Burton (Eds) Geology of the National Capital Region– Field Trip Guidebook, US Geological Survey Circular, v. 1264, p. 77-135. Cecil, C. B., DiMichele, William, Fedorko, Nick, and Skema, Vik, 2011, Autocyclic and allocyclic controls on the origin of the , in J. A. Harper, (Ed), Geology of the Pennsylvanian-Permian in the Dunkard basin. Guidebook, 76th Annual Field Conference of Pennsylvania Geologists, p. 26-45. Cecil, C.B., Dimichele, W.D., and Elrick, S.D., 2014, Middle and Late Pennsylvanian cyclothems, American midcontinent: Ice-age environmental changes and terrestrial biotic dynamics, External Geophysics, Climate, Comptes Rendus Geoscience, Elsevier, v. 346, p. 159-168. Chesnut, D.R. Jr., 1994, Eustatic and tectonic control of deposition of the Lower and Middle Pennsylvanian strata of the central Appalachian Basin, in Tectonic and eustatic controls on sedimentary cycles, Society for Sedimentary Geology, Concepts in Sedimentology and Paleontology, v. 4, p. 51–64. Christensen, A.M., 1999, Brachiopod paleontology and paleoecology of the Lower Lodgepole Limestone in Southeastern Idaho, in Hughes, S.S., and Thackray, G.D., (Eds), Guidebook to the Geology of Eastern Idaho, Pocatello, Idaho Museum of Natural History, p. 57-67. Collinson, J.D., 1986, Alluvial sediments, in, H.G. Reading (Ed), Sedimentary Environments and Facies (2nd ed.), Oxford, Blackwell Scientific Publications, p. 20-62. DeCelles, P. G., and Giles, K. N., 1996, Foreland basin systems, Basin Research, v. 8, p.105-123. 109

DeCelles, P., Kapp, P., Quade, J., and Gehrels, G., 2011, Kailas basin, southwestern Tibet: Record of postcollisional upperplate extension in the Indus-Yarlung suture zone, Geological Society of America Bulletin, v. 123, p. 1337–1362. Dennison, J. M., 1989, Paleozoic sea-level changes in the Appalachian Basin: Washington, D.C., 28th International Geological Congress Field Trip Guidebook, American Geophysical Union Field Trip, T354, p. 50-56. Dennison, J.M., and Ettensohn, F.R., 1994, Tectonic and eustatic controls on sedimentary cycles, SEPM Concepts in Sedimentology and Paleontology, v. 4, p. 264. DiMichele, W.A., Cecil, C.B., Montanez, I.P., and Falcon-Lang, H.J., 2010, Cyclic changes in Pennsylvanian paleoclimate and effects on floristic dynamics in tropical Pangaea, International Journal of Coal Geology, v. 83, p. 329-344. Donaldson, A. C., 1979, Depositional environments of the upper Pennsylvanian series, in, K.J. Englund, H.H. Arndt, and T.W. Henry (Eds), Proposed Pennsylvanian System Stratotype Virginia and , American Geologic Institute, Selected Guidebook Series 1, p. 123–132. Donaldson, A. C., Renton J. J., and Presley M. W., 1985, Pennsylvanian deposystems and paleoclimates of the Appalachians, International Journal of Coal Geology, v. 5, p. 167–193. Driese, S.G., and Ober, E.G., 2005, Paleopedologic and paleohydrologic records precipitation seasonality from Early Pennsylvanian “Underclay” paleosols, U.S.A., Journal of Sedimentary Research, v. 75, p. 997-1010. Dyni, J.R., 2006, Geology and resources of some world oil-shale deposits, U.S. Geological Survey Scientific Investigations Report 2005–5294, p. 42. Eble, C.F., 2002, Palynology of late Middle Pennsylvanian coal beds in the Appalachian Basin, International Journal of Coal Geology, v. 50, p. 73-88. Edmunds, W.E., Berg, T.M., Sevon, W.D., Piotrowski, R.C., Hayman, L., and Rickard, L.V., 1979, Pennsylvania and New York, in the Mississippian and Pennsylvanian (Carboniferous) systems in the , U.S. Geological Survey Professional Paper 1110–B, p. B1–B33. Ettensohn, F. R., 1994, Tectonic control on the formation and cyclicity of major Appalachian unconformities and associated stratigraphic sequences, in, J.M. Dennison and F.R. Ettensohn (Eds), Tectonic and eustatic controls on sedimentary cycles, SEPM Concepts in Sedimentology and Paleontology, v. 4, p. 217-242. Ferm, J. C., 1970, Allegheny deltaic deposits, in, J.P. Morgan and R.H. Shaver (Eds), Deltaic Sedimentation; Modern and Ancient, SEPM Special Publication, v. 15, p. 246-255. 110

Fischer, A. G., 1982, The two Phanerozoic supercycles, in, W.A. Berggren and J.A. van Couvering (Eds), Catastrophes and Earth History, Princeton Press, p. 129-150. Flemming, E. A., 2003, Incised valley fill: a sequence stratigraphic and depositional model for the Lower Freeport Sandstone, Athens and Vinton Counties, Ohio, Master’s Thesis, Ohio University, Athens, Ohio, p. 81. Flores, R. M., 1968, Simple linear regression as a means of predicting the thickness of Middle Allegheny stratigraphical interval in eastern Ohio, Journal of Sedimentary Petrology, v. 38, p. 400-410. Flores, R. M., 1979, Relationship of geometry to the origin of the Lower Freeport Sandstone (Middle Allegheny) of eastern Ohio, in, J.C. Ferm, J.C. Horne, J.R. Staub, and G. A. Weisenfluh (Eds), Carboniferous Depositional Environments in the Appalachian Region, University of South Carolina Department of Geology, Carolina Coal Group, p. 147-154. Folk, R.L., 1951, Stages of textural maturity in sedimentary rocks, Journal of Sedimentary Petrology, v. 21 p. 127-130. Folk, R.L., Andrews P.B., and Lewis, D.W., 1970, Detrital sedimentary rock classification and nomenclature for use in New Zealand, New Zealand Journal of Geology and Geophysics v. 13 p. 937-968. Galehouse, J. S., 1969, Counting of grain mounts number percentage vs number frequency: Journal of Sedimentary Petrology, v. 39, p. 812-815. Gardner, T. W., Williams E. G., and Holbrook, P. W.,1988, Pedogenesis of some Pennsylvanian underclays; Ground-water, topographic, and tectonic controls, in, J. Reinhardt and W.R. Sigleo (Eds), Paleosols and Weathering Through Geologic Time, Geological Society of America, Special Paper 216, p. 81-100. Gastaldo, R. A., Purkynova, E., Simunek, Z., and Schmitz, M.D., 2009, Ecological persistence in the Late Mississippian (- A) megafloral record of the Upper basin, Czech Republic, Palaios, v. 24, p. 336–350. Gautier, D.L., 1982, Siderite concretions—Indicators of early diagenesis in the Gammon Shale (), Journal of Sedimentary Petrology, v. 52, p. 859-871. Greb, S.F., and Chestnut, D. R., Jr., 1996, Lower and lower Middle Pennsylvanian fluvial to estuarine deposition, central Appalachian basin: Effects of eustacy, tectonics, and climate, Geological Society of America Bulletin, v. 108, p. 303-317. Greb, S.F., Andrews, W.M., Eble, C.F., DiMichele, W.A., Cecil, C.B., and Hower, J.C., 2003, Desmoinesian coal beds of the Eastern Interior and surrounding basins: the largest tropical peat mires in Earth history, in, M.A. Chan and A.W. Archer (Eds), Extreme Depositional Environments: Mega End Members in Geologic Time, Geological Society of America Special Paper 370, p. 127–150. Greb, S. F., Pashin, J. C., Martino, R. L., and Eble, C. F., 2008, Appalachian sedimentary cycles during the Pennsylvanian: Changing influences of sea level, climate, and tectonics, in, C.R. Fielding, T.D. Frank, and J.L. Isbell (Eds), 111

Resolving the Late Paleozoic Ice Age in Time and Space, Geological Society of America, Special Paper 441, p. 235-248. Harms, J. C., Southard J. B, and Walker R. G., 1982, Structures and sequences in clastic rocks, SEPM Short Course No. 9, p. 3-21. Heckel, P. H., 1977, Origin of phosphatic black shale facies in Pennsylvanian cyclothems of mid-continent North America, Bulletin of the American Association of Petroleum Geologists, v. 61, p. 1045-1068. Heckel, 1995, Glacial-eustatic Base level-climate model for late Middle to Late Pennsyvanian coal-bed Formation in the Appalachian Basin, Journal of Sedimentary Research, B65, p. 348-356. Heckel, P. H. 1989, Updated Middle-Upper Pennsylvanian Eustatic Sea Level Curve for Midcontinent North America and Preliminary Biostratigraphic Characterization, Onzieme Congres Intemational de Stratigraphic et de Geologic du Carbonifere, Compte Rendu, v. 4, p. 160-185. Heckel, P.H., 2003, Updated cyclothem constraints on radiometric dating of the Pennsylvanian succession in North America and its correlation with dates from Europe, Newsletter on Carboniferous Stratigraphy, v. 21, p. 12-20. Heckel, P. H., 2008, Pennsylvanian cyclothems in midcontinent North America as far- field effects of waxing and waning of Gondwana ice sheets in, C.R. Fielding, T.D. Frank, and J.L. Isbell (Eds), Resolving the Late Paleozoic Ice Age in Time and Space, Geological Society of America, Special Paper 441, p. 275–289. Heckel, P.H., 2013, Pennsylvanian stratigraphy of Northern Midcontinent Shelf and biostratigraphic correlation of cyclothems, Stratigraphy, v. 10, p. 3-39. Heckel, P.H., Alekseev, A.S., Barrick, J.E., Boardman, D.R., Goreva, N.V., Isakova, T.I., Nemyrovska, T.I., Ueno, K., Villa, E. and Work, D.M., 2008, Choice of conodont simulator (sensu stricto) as the event marker for the base of the global Gzhelian Stage (Upper Pennsylvanian Series, Carboniferous System), Episodes, v. 31, p. 319-325. Hembree, D. I., and Nadon, G. C., 2011, A paleopedologic and ichnologic perspective of the terrestrial Pennsylvanian landscape in the distal Appalachian Basin, USA. Palaeogeography, Palaeoclimatology, Palaeoecology, v. 312, p.138-166. Horton, R.E., 1945, Erosional development of streams and their drainage basins: hydrophysical approach to quantitative morphology, Geological Society of America, Bulletin, v. 56, p. 275–370. Hughes, R.E., Demaris, P.J., and White, W.A., 1992, Underclays and related paleosols associated with coals, in, I.P. Martini and W. Chesworth (Eds), Developments in Earth Surface Processes 2, Weathering and Paleosols. Amsterdam, Elsevier, p. 501-523. Jervey, M. T., 1988, Quantitative geological modeling of siliciclastic rock sequences and their seismic expression, in, C.K. Wilgus, B.S. Hastings, C.G.St.C Kendall, H.W. 112

Posamentier, C.A. Ross, and J.C. Van Wagoner (Eds), Sea Level Research: an Integrated Approach, SEPM Special Publication no. 42, p. 47-69. Joeckel, R. M., 1995, Paleosols below the Ames marine unit (Upper Pennsylvanian Conemaugh Group) in the Appalachian Basin, USA: variability on an ancient depositional landscape, Journal of Sedimentary Research, v. 65, p. 393–407. Johnson, T. A., 2004, Stratigraphy, depositional environments, and coalbed gas potential of Middle Pennsylvanian (Desmoinesian Stage) coal--Bourbon arch region, eastern Kansas, Kansas Geological Survey, Open-file Report 2004-38, p. 309. Kamola, D. L., and Van Wagoner, J. C., 1995, Stratigraphy and facies architecture of parasequences with examples from the Spring Canyon Member, Blackhawk Formation, Utah, in, J.C. Van Wagoner and C.T. Bertram (Eds), Sequence Stratigraphy of Foreland Basin Deposits, Outcrop and Subsurface Examples from the Cretaceous of North America, American Association of Petroleum Geologists Memoir 64, p. 27-54. Kosanke, R. M., 1988, Palynological studies of Middle Pennsylvanian coal beds of the proposed Pennsylvanian System stratotype in West Virginia, U.S. Geological Survey, Professional Paper 1455, p. 73. Kosanke, R.M., and Cecil, C.B., 1996, Late Pennsylvanian climate changes and palynomorph extinctions, Review of Palaeobotony and Palynology, v. 90, p. 113- 140. Kozur, H. W., and Weems, R. E., 2010, The biostratigraphic importance of conchostracans in the continental of the northern hemisphere, in, S.G. Lucas (Ed), The Triassic Timescale, Geological Society, London, Special Publications 334, p. 315– 417. Kraus, M.J., 1999, Paleosols in clastic sedimentary rocks: their geologic applications, Earth Science Reviews, v. 47, p. 41-70. Kraus, M.J., and Hasiotis, S.T., 2006, Significance of different modes of rhizolith preservation to interpreting paleoenvironmental and paleohydrologic settings: examples from Paleogene paleosols, Bighorn Basin, Wyoming, USA, Journal of Sedimentary Research, v. 76, p. 633-646. Logan, W.E., 1842, On the character of the beds of clay lying immediately below the coal seams of south Wales, Geological Society of London, Proceedings, v. 3, p. 275–277. Loutit, T. S., Hardenbol, J., Vail, P. R., and Baum, G. R., 1988, Condensed sections: the key to age determination and correlation of continental margin sequences, in, C.K. Wilgus, B.S. Hastings, C.G.St.C Kendall, H.W. Posamentier, C.A. Ross, and J.C. Van Wagoner (Eds), Sea Level Research: an Integrated Approach, SEPM Special Publication no. 42, p.183-213. Liu, Y., Gastaldo, R.A., 1992, Characteristics of a Pennsylvanian ravinement surface, Sedimentary Geology, v. 77, p. 197-208. 113

Martino, R.L., 2016, Sequence stratigraphy of fluvially dominated strata of the Mid-Late Pennsylvanian Conemaugh Formation, Central Appalachian Basin, International Journal of Coal Geology, v. 154-155, p. 136-154. MacEachern, J. A., Raychaudhuri, I., and Pemberton, S. G., 1992, Stratigraphic applications of the Glossifungites Ichnofacies—Delineating discontinuities in the rock record; in, S. G. Pemberton (Ed.) Applications of Ichnology to Petroleum Exploration—A Core Workshop, SEPM Core Workshop 17, p. 169-198. Mack, G. H., James, H.C., and Monger, H.C., 1993, Classification of paleosols, Geological Society of America Bulletin, v. 105, p. 129-136. McCabe, P.J., 1984, Depositional models of coal and coal-bearing strata; in, Sedimentology of coal and coal-bearing sequences, in, R.A. Rahmani and R.M. Flores (Eds), Coal and Coal-bearing Sequences, International Association of Sedimentologists, Special Publication 7, p. 13-42. Menning, M., Alekseev, A. S., Chuvashov, B. I., Davydov, V. I., Devuyst, F. -X., Forke, H. C., Grunt, T. A., Hance, L., Heckel, P. H., Izokh, N. G., Jin, Y. -G., Jones, P. J., Kotlyar, G. V., Kozur, H. W., Nemyrovska, T. I., Schneider, J. W., Wang, X. - D., Weddige, K., Weyer, D., and Work, D. M., 2006, Global time scale and regional stratigraphic reference scales of Central and West Europe, East Europe, Tethys, South China, and North America as used in the – Carboniferous–Permian correlation chart 2003 (DCP 2003), Palaeogeography, Palaeoclimatology, Palaeoecology, v. 240, p. 318–372. Merrill, W. M. 1950. The Geology of Northern Hocking County, Ohio* Unpublished Dissertation, The Ohio State University, p. 444. Miall A.D., 1977, A review of the braided river depositional environment, Earth Science Reviews., v. 13, p. 1-62. Miall, A.D., 2013, Fluvial Depositional Systems, Springer-Verlag, p. 600. Miall, A.D., 2016. Facies models. In: Miall, A.D. (Ed), Stratigraphy: A Modern Synthesis. Springer International Publishing, p. 161-214. Milici, R.C., Freeman, P.A, Carlton, R.W., Bragg, L.J., Butler, D.T., and Tewalt, S.J., 2001, Chapter E, Geology and geochemistry of the Middle Pennsylvanian Lower Kittanning coal bed, Allegheny Group, northern Appalachian coal region, U.S. Geological Survey Professional Paper 1625-C, p. E1 – E67. Mitchum, R.M., 1977, Seismic stratigraphy and global changes of sea level, part 1: Glossary of terms used in seismic stratigraphy, in, C.E. Payton (Ed), Seismic Stratigraphy— Applications to Hydrocarbon Exploration, American Association Petroleum Geologists Memoir 26, p. 205-212. Nadon, G. C., 1994, The genesis and recognition of anastomosed fluvial deposits: Data from the St. Mary River Formation, southwestern Alberta, Canada, Journal of Sedimentary Research, Section B: Stratigraphy and Global Studies, v. 64B(4), p. 451-463. 114

Nadon, G.C. and Hembree, D.I., 2007, Variations in sequence, fluvial, and pedogenic architecture in the distal Appalachian foreland basin, SE Ohio, SEPM Great Lakes Section Field Conference Guidebook, p. 1-42. Opdyke, N.D. and DiVenere, V.J., 1994, Paleomagnetism and Carboniferous climate, in, C.B. Cecil and N.T. Edgar (Eds), Predictive Stratigraphic Analysis - Concept and Application, United States Geological Survey Bulletin 2110, p. 8-10. Plint, A.G., and Nummedal, D., 2000, The falling stage systems tract: recognition and importance in sequence stratigraphic analysis; in, D. Hunt and R. Gawthorpe (Eds) Sedimentary Responses to Forced Regressions, Geological Society of London, Special Publication 172, p. 1-17. Plafker, G., and Sasvage, J. C., 1970, Mechanism of the Chilean earthquake of May 21 and 22, I960, Geological Society of America Bulletin, v. 81, p. 1001-1030. Potsma, D., 1982, Pyrite and siderite formation in brackish and freshwater swamp sediments, American Journal of Science, v. 282, p. 1151-1183. Pearson, N.J., Mángano, M.G., Buatois, L.A., Casadío, S., and Rodriguez, M., 2012, Ichnology, sedimentology, and sequence stratigraphy of outer-estuarine and coastal-plain deposits: Implications for the distinction between allogenic and autogenic expressions of the Glossifungites Ichnofacies, Palaeogeography, Palaeoclimatology, Palaeoecology, v. 333, p. 192-217. Pemberton, S. G., and Wightman, D. M., 1992, Ichnological characteristics of brackish water deposits; in, S. G. Pemberton (Ed) Applications of Ichnology to Petroleum Exploration—A Core Workshop, SEPM Core Workshop 17, p. 141–167. Quinlan, G. M., and Beaumont, C., 1984, Appalachian thrusting, lithospheric flexure, and the Paleozoic stratigraphy of the eastern interior of North America, Canadian Journal of Earth Sciences, v. 21, p. 973–996. Retallack, G.J., 1988, Field Recognition of paleosols, in, J. Reinhardt and W.R. Sigleo (Eds), Paleosols and Weathering Through Geologic Time, Geological Society of America, Special Paper 216, p. 1-20. Retallack, G.J., 1990, Soils of The Past, Unwin-Hyman, p. 520. Ross, C.A. and Ross, J.R., 1987, Late Paleozoic sea levels and depositional sequences, in, C.A. Ross and D. Haman, (Eds), Timing and Depositional History of Eustatic Sequences; Constraints on Seismic Stratigraphy: Cushman Foundation for Foraminiferal Research, Special Publication, v. 24, p. 137-149. Ross, C. A., and J. R. P. Ross., 1988, Late Paleozoic transgressive-regressive deposition, in, C.K. Wilgus, B.S. Hastings, C.G.St.C Kendall, H.W. Posamentier, C.A. Ross, and J.C. Van Wagoner (Eds), Sea Level Research: an Integrated Approach, SEPM Special Publication no. 42, p. 227-247. Rubin D, M., and Carter C. L, 1987, Cross-bedding, Bedforms, and Paleocurrents, SEPM Concepts in Sedimentology and Paleontology, v. 1, p. 187. 115

Schieber J., 2011, Shales, in, J. Reitner and V. Thiel (Eds), Encyclopedia of Giobiology, Springer Verlag, p. 785-791. Schumm, S. A., 1981, Evolution and response of the fluvial system, sedimentologic implications, SEPM Special Publication 31, p. 19-29. Scotese, C.S., 2003, Paleomap Project, Web site accessed May 4, 2016, http://www.scotese.com. Stout, W., 1923, Origin of coal formation clays, Geological Survey of Ohio, Fourth Series, Bulletin, v. 26, p. 533–568. Stout, W., 1939, Generalized Section of Coal Bearing Rocks of Ohio, Geological Survey of Ohio, 4th ser., Information Circular, No. 2, p. 6. Stout, W., 1947, Generalized Section of Rocks of Ohio, Geological Survey of Ohio, 4th ser., Information Circular, No. 4, p. 108. Sturgeon, M.T., 1958, The geology and mineral resources of Athens County, Ohio Department of Natural Resources, Division of Geological Survey, Bulletin, v. 57, p. 43-591. Sturgeon, M.T., and Merrill, W. M., 1949, An additional fossiliferous member in the (Pennsylvanian) of Ohio, The Ohio Journal of Science, v. 49, p. 1-11. Tankard, A.J., 1986, On the depositional response to thrusting and lithospheric flexure: Examples from the Appalachian and Rocky Mountain basins, in, P.A. Allen and P. Homewood (Eds), Foreland Basins, International Association of Sedimentologists Special Publication 8, p. 369-394. Thomas, W.A., 1976, Evolution of Ouachita-Appalachian continental margin, Journal of Geology, v. 84, p. 323-342. Udden, J.A., 1912, Geology and mineral resources of the Peoria quadrangle, Illinois, U.S. Geological Survey, Bulletin, v. 506, p. 47–50. Van Wagoner, J. C., Mitchum, R. M., Campion, K. M., and Rahmanian, V. D., 1990, Siliciclastic Sequence Stratigraphy in Well Logs, Cores, and Outcrops: Concepts for High-resolution Correlation of Time and Facies, American Association of Petroleum Geologists Methods in Exploration Series, no. 7, p. 1-30. Van Wagoner, J. C., Posamentier, H. W., Mitchum, R. M. J., Vail, P. R., Sarg, J. F., Loutit, T. S., and Hardenbol, J., 1988, An overview of the fundamentals of sequence stratigraphy and key definitions, in, C. K. Wilgus, B. S. Hastings, C. G. St. C. Kendall, H. W. Posamentier, C. A. Ross, and J. C. Van Wagoner (Eds), Sea-level changes: An integrated approach, SEPM Special Publication, no. 42, p. 39–45. Van Wagoner, J.C., and G.T. Bertram, (Eds), 1995, Sequence Stratigraphy of Foreland Basin Deposits - Outcrop and Subsurface Examples from the Cretaceous of North America, American Association of Petroleum Geologists Memoir 64, p. 489. 116

Vail, P. R., Mitchum, R.M., Todd, R.G., Widmier, J.M., Thompson, S., Sangree, J.B., Bubb, J.N., and Hatfield, W.G., 1977, Seismic stratigraphy and global changes in sea level, in, C.E. Payton (Eds.), Seismic stratigraphy— Applications to hydrocarbon exploration, American Association of Petroleum Geologists Memoir 26, p. 49–212. Walker, R. G., 1976, Facies models 3, Sandy Fluvial Systems, Geoscience Canada, v. 3 (2), p. 101-109. Walker, R.G., and Cant, D.J., 1984, Sandy fluvial systems, in, R.G Walker (Ed), Facies Models, 2nd ed., Geoscience, Canada, p. 71-89. Wanless, H.R., 1931, Pennsylvanian cycles in western Illinois, Illinois State Geological Survey, Bulletin, v. 60, p. 179-193. Wanless, H.R. and Weller, J.M., 1932, Correlation and extent of Pennsylvanian cyclothems, Geological Society of America, Bulletin, v. 43, p. 1003-1016. Watts, A.B., Karaer, G. D., and Steckler, M.S., 1982, Lithospheric flexture and the evolution of sedimentary basins, Royal Society of London Philosophical Transactions, Series A, v. 305, p. 249-281. Wright, W. F., 1979, Petroleum Geology of the Permian Basin, Midland, Texas, West Texas Geological Society, p. 98. Zecchin, M., Caffau, M., Catuneanu, O., and Lenaz, D., 2017, Discrimination between wave-ravinement surfaces and bedset boundaries in shallow-marine deposits, Crotone Basin, southern Italy: An integrated sedimentological, micropalaeontological and mineralogical approach, Sedimentology, v. 64, p. 1756. Ziegler, A.M., Raymond, A.L., Gierlowski, T.C., Horell, M.A., Rowley, D.B., and Lottes, A.L., 1987, Coal climate and terrestrial productivity: The present and compared, in, A.C. Scott (Ed), Coal and Coal Bearing Strata: Recent Advances, Geological Society of London, Special Publication 32, p. 25-49.

117

APPENDIX 1: DETAILED MEASURED SECTIONS

118

119

120

121

122

123

APPENDIX 2: PALEOCURRENT DIRECTIONS

Nelsonville Bypass Paleocurrent Data

Outcrop 39.471786, -82.251044 Location (OC-3)

No. Strike Dip Direction No. Strike Dip Direction 1 150 15 W 26 180 24 W 2 184 6 W 27 150 17 W 3 198 12 W 28 110 20 N 4 110 15 S 29 140 9 W 5 32 22 N 30 15 36 W 6 83 12 N 31 90 20 N 7 55 20 N 32 192 15 W 8 80 19 N 33 105 29 N 9 135 5 W 34 30 15 N 10 40 13 N 35 105 23 N 11 40 6 N 36 51 19 N 12 45 15 N 37 100 10 S 13 50 15 N 38 63 21 N 14 25 20 W 39 42 15 N 15 59 24 N 40 40 20 N 16 4 11 E 41 200 9 W 17 85 20 N 42 110 25 N 18 58 8 N 43 205 6 E 19 199 14 W 44 105 31 N 20 170 22 W 45 100 25 N 21 180 17 W 46 145 10 E 22 190 16 W 47 12 34 W 23 127 25 S 48 110 29 N 24 152 15 E 49 105 8 N 25 184 22 W 50 10 28 W

124

APPENDIX 3: POINT COUNTING

OC1-B-St OC1-M-St OC5-MK-St OC3-RLB1 OC5-RLB1 Monocrystalline 292 304 214 138 69 Quartz Polycrystalline 159 115 110 53 45 Quartz Pore 0 2 4 1 0 Na-Ca Feldspar 32 28 27 8 6 Biotite 1 49 11 13 1 Muscovite 24 37 13 27 12 Limonite 0 0 13 2 1 K-Feldspar (Orth) 0 0 0 39 27 Garnet 9 1 1 0 0 Opaque 5 4 2 4 5 Quartz Matrix 78 60 29 37 4 Calcite Matrix 0 0 176 264 428 Unknown 0 0 0 14 0 Zircon 0 0 0 0 2 Total 600 600 600 600 600

125

APPENDIX 4: XRD DATA

XRD Sample Preparation

Samples submitted for XRD analysis are first disaggregated using a mortar and pestle, weighed, and dispersed in de-ionized water using a sonic probe. The samples are next centrifugally size fractionated into a bulk (4 microns) and a clay-size (<4 microns ESD) fraction. The clay suspensions are then decanted and vacuum-deposited on nylon membrane filters to produce oriented mounts.

Clay mounts are attached to glass slides and exposed to ethylene glycol vapor for a minimum of 24 hours to aid in detection and characterization of expandable clays. The bulk fractions of each sample are dried and weighed in order to determine weight loss due to removal of clay-size materials.

Analytical Procedures

XRD analyses of the clay-size fractions of the samples are performed using a

Siemens D500 automated powder diffractometer equipped with a CuKa radiation source (40 Kv, 35 mA) and a solid state or scintillation detector. The air-dried and glycol-solvated oriented clay mounts are analyzed over an angular range of

2-36 degrees 2 theta at a scan rate of 1 degree/minute. Quantitative analyses of the diffraction data are done using integrated peak areas (derived from peak deconvolution / profile-fitting techniques) and empirical reference intensity ratio

(RIR) factors determined specifically for the diffractometer used for data collection. Determinations of mixed-layer clay type, ordering and percent expandable interlayers are done by comparing experimental diffraction data from 126 the glycol-solvated clay aggregates with simulated one dimensional diffraction profiles generated using the program NEWMOD written by R. C. Reynolds.

127

Brookville XRD Data (wt%) TOTAL (wt%) Jarosite e(wt%) Dolomit (wt%) Calcite 0.00.00.00.0 0.0 0.00.0 0.0 0.00.0 0.0 0.00.0 0.00.0 0.0 0.0 0.0 0.0 8.6 0.0 0.0 0.0 100.0 0.0 2.9 0.0 100.0 0.0 100.0 0.0 0.0 100.0 0.0 0.0 100.0 0.0 100.0 100.0 100.0 (wt%) Plagioclase (wt%) K-Feldspar (wt%) Quartz Total Clay (%) Clay (wt%) Chlorite (wt%) Kaolinite 3.933.402.231.56 93.4 86.1 6.44 92.8 2.81 86.5 2.201.77 93.7 94.4 88.8 90.4 Kaolinite/Mica 17.718.427.131.9 69.5 62.611.7 60.3 3.423.4 49.7 3.626.1 3.4 5.030.7 75.3 3.7 4.1 65.8 6.3 57.4 5.1 10.0 1.6 54.2 3.2 1.2 4.2 0.9 4.9 0.6 3.5 4.7 10.3 1.4 9.2 0.9 0.9 0.4 (wt%) IlliteMica& (wt%) >4 micron >4 2.8 1.5 2.0 1.2 1.6 2.0 1.1 2.0 20S* R1 M-L R1 I/S 8.99.79.6 91.1 90.3 8.7 90.4 7.27.7 91.3 92.8 92.3 20.8 79.2 15.4 84.6 (wt%) <4 micron <4 0 0 25 50 75 25 50 75 Depth(cm) Outcrop# Outcrop# XRD XRD XRD XRD UN137UN135 OC7-RLB2-1W UN133 OC7-RLB2-2W UN131 OC7-RLB2-3W OC7-RLB2-4W UN138UN136 OC7-RLB2-1F UN134 OC7-RLB2-2F UN132 OC7-RLB2-3F OC7-RLB2-4F UN137 OC7-RLB2-1W UN135UN133 OC7-RLB2-2W UN131 OC7-RLB2-3W OC7-RLB2-4W UN138UN136 OC7-RLB2-1F UN134 OC7-RLB2-2F UN132 OC7-RLB2-3F OC7-RLB2-4F sample# sample# 128

Lower Kittanning XRD Data (wt%) T OT A L (wt%) Jaro site (wt%) D o lo mite 0.0 0.0 0.0 100.0 0.00.00.00.0 0.00.0 0.00.0 0.00.0 0.0 7.40.0 0.0 5.7 0.0 0.00.0 0.0 100.0 0.00.0 0.0 100.0 0.00.0 100.0 0.00.0 0.0 100.0 1.00.0 0.0 100.0 0.00.0 0.0 100.0 0.0 0.0 100.0 0.0 0.0 100.0 0.0 0.0 0.0 0.0 100.0 0.0 100.0 0.0 100.0 0.0 100.0 0.0 100.0 100.0 100.0 (wt%) C alcite 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 0.0 (wt%) Plagioclase 1.2 0.5 0.3 0.6 0.3 0.5 0.8 0.6 0.5 0.4 0.4 0.6 0.6 (wt%) K-Feldspar (wt%) Quartz 91.5 95.1 88.2 96.3 94.3 92.5 88.9 93.0 90.2 88.4 85.7 95.0 95.6 94.8 95.4 94.0 T o tal C lay (%) 2.1 3.9 3.1 4.9 3.5 3.1 0.6 3.2 9.3 3.63.33.5 8.0 4.7 3.2 11.3 3.44.0 5.9 3.6 10.43.62.9 3.9 4.6 4.8 0.7 2.7 4.2 2.8 8.1 4.8 4.3 6.4 0.4 (wt%) C hlo rite 1.94 3.61 4.14 11.52 5.08 3.08 3.99 7.67 4.40 3.80 7.93 7.76 3.98 2.98 3.82 16.25 84.1 63.1 71.2 71.9 71.2 68.3 74.4 64.7 70.3 53.6 74.7 68.6 79.3 79.2 64.3 82.9 (wt%) Kaolinite/M ica Kao linite 86.1 78.1 87.1 95.0 77.7 86.2 86.9 84.7 82.6 94.8 95.7 78.2 89.0 90.9 92.9 87.5 (wt%) 5.1 7.3 17.1 9.7 14.7 18.5 14.7 19.0 10.0 10.2 17.9 21.6 18.8 17.2 27.6 20.5 >4 micro n (wt%) Illite & M ica 9.1 5.2 4.3 11.0 21.8 21.9 (wt%) 1.8 2.2 2.7 2.9 3.5 2.3 2.7 2.7 20S* <4 micro n R1 M -L I/S 25 50 75 25 50 75 (cm) 25 50 75 25 50 75 D epth (cm) D epth Outcro p Sample # Outcro p Sample # XR D XR D Sample # Sample # UN153 UN155 UN157 OC4-LKC-5WUN159 OC4-LKC-6W OC4-LKC-7W UN146 OC4-LKC-8W 100 125 OC4-LKC-1F 150 175 5.0 22.3 13.8 0 13.1 UN160 OC4-LKC-8F 13.9 175 16.3 UN145 UN145 OC4-LKC-1W 0 17.4 UN147 UN147 UN149 UN151 OC4-LKC-2W OC4-LKC-3W OC4-LKC-4W UN148 UN150 UN152 OC4-LKC-2F UN154 OC4-LKC-3F UN156 OC4-LKC-4F UN158 OC4-LKC-5F OC4-LKC-6F OC4-LKC-7F 100 125 150 7.1 12.5 12.9 UN145 UN145 UN147 UN149 OC4-LKC-1W UN151 OC4-LKC-2W UN153 OC4-LKC-3W UN155 OC4-LKC-4W 0 UN157 OC4-LKC-5WUN159 OC4-LKC-6W OC4-LKC-7W UN146 OC4-LKC-8W 100 UN148 125 UN150 OC4-LKC-1F 150 UN152 OC4-LKC-2F 175 1.4 UN154 OC4-LKC-3F 1.8 UN156 OC4-LKC-4F 0 1.7 UN158 OC4-LKC-5F 1.9 UN160 OC4-LKC-6F OC4-LKC-7F OC4-LKC-8F 100 125 150 175 1.4 1.7 1.8 1.6

129

Upper Freeport Paleosol XRD Data (wt%) T OT A L 0.02.9 100.0 3.6 100.0 0.00.0 100.0 100.0 100.0 (wt%) Jaro site (wt%) D o lo mite 0.00.00.0 0.0 0.0 0.0 0.0 0.00.0 0.0 0.0 12.2 0.0 100.0 (wt%) C alcite 0.0 0.0 0.0 0.0 0.0 0.0 (wt%) Plagioclase 1.0 1.4 1.2 0.6 0.6 0.7 (wt%) K-Feldspar (wt%) Quartz 5.1 9.4 4.54.04.4 10.9 3.0 8.1 4.3 5.7 4.9 9.6 (wt%) C hlo rite 90.1 78.7 94.4 89.0 85.6 (%) 89.4 Total Clay 64.1 60.8 52.8 45.0 60.7 47.9 (wt%) Kao linite 2.11 1.65 3.19 1.38 2.03 2.76 20.1 22.0 26.0 27.3 28.7 34.6 (wt%) Kaolinite/M ica Illite & M ica 96.7 94.4 98.7 (wt%) 1.5 1.5 2.3 2.4 2.0 2.2 20S* >4 micro n R1 M -L I/S (wt%) (cm) D epth <4 micro n Outcro p Sample # Outcro p Sample # UN141 0C4-LFC-2W 1.3 UN141 0C4-LFC-2W 25.0 UN140UN142 0C4-LFC-1F 0C4-LFC-2F 3.3 5.6 UN143 0C4-LFC-3W 11.4 88.6 XR D UN139 0C4-LFC-1W 12.9 87.1 UN143 0C4-LFC-3WUN140UN142 0C4-LFC-1F 50.0 0C4-LFC-2F 0.0 25.0 XR D UN139 0C4-LFC-1W 0.0 UN144 UN144 0C4-LFC-3F 11.0 89.0 UN144 UN144 0C4-LFC-3F 50.0 Sample # Sample #

130

APPENDIX 5: XRF DATA Total mass% P2O5 mass% K2O mass% Na2O mass% CaO mass% MnO mass% MgO mass% Fe2O3 mass% Al2O3 mass% TiO2 mass% SiO2 mass% Date / Time / Date Analysis Procedure Analysis Outcrop Outcrop Sample # Sample 0C4-LFC-1F0C4-LFC-2W0C4-LFC-2F1c App Dil7-1 FP 1c App Dil7-1 FP 0C4-LFC-3W10:46 2/21/2017 0C4-LFC-3F1c App 13:49 Dil7-1 FP 2/21/2017 55.621c App Dil7-1 FP OC4-LKC-1W 55.5414:19 2/21/2017 OC4-LKC-1F1c App 14:50 Dil7-1 FP 1.20 2/21/2017 1c 55.37App Dil7-1 FP 1.19OC4-LKC-2W 56.7015:20 2/21/2017 1c 24.59App 15:03 Dil7-1 OC4-LKC-2F FP 3/17/2017 1.041c 25.17 55.93App Dil7-1 FP 1.20OC4-LKC-3W 50.0716:21 2/21/2017 3.671c 25.12App 16:52 Dil7-1 OC4-LKC-3F FP 2/21/2017 3.00 1.07 53.231c 25.06App Dil7-1 FP 1.01OC4-LKC-4W 48.5519:55 0.85 2/21/2017 3.881c 24.59App 20:25 Dil7-1 OC4-LKC-4F FP 1.05 0.80 2/21/2017 2.99 26.95 53.761c App Dil7-1 FP 0.98OC4-LKC-5W 0.02 55.40 1.048:57 4.09 2/22/2017 29.231c 0.02App 1.99Dil7-1 OC4-LKC-5F FP 1.09 0.849:28 2/22/2017 26.151c App Dil7-1 FP 1.04 54.59 0.21OC4-LKC-6W 0.06 1.69 1.089:58 2/22/2017 56.28 29.20 0.10 0.411c 0.02App 10:29 5.97Dil7-1 OC4-LKC-6F FP 2/22/2017 27.37 1.071c App Dil7-1 FP 0.12 56.95 0.20OC4-LKC-7W 0.43 0.04 64.0510:59 1.01 1.69 2/22/2017 0.20 0.01 0.06 0.481c App 11:30 2.06Dil7-1 OC4-LKC-7F FP 27.98 2/22/2017 65.08 1.131c 2.82App Dil7-1 FP 1.92 0.18 26.95 0.01 0.35OC4-LKC-8W 0.45 62.5212:00 3.10 2/22/2017 0.03 0.33 0.02 1.99 0.701c App 12:31 Dil7-1 OC4-LKC-8F FP 2.18 26.70 2/22/2017 21.67 1.67 61.81 0.071c 3.52App Dil7-1 0.05 FP 1.34 0.22 0.01 62.87 0.0813:01 0.13 3.61 2/22/2017 0.05 0.55 0.01 20.43 1.661c App 13:32 1.23Dil7-1 FP 1.23 89.16 2/22/2017 0.69 23.43 62.85 0.11 0.06 3.78 0.07 1.18 89.19 1.67 60.74 0.0914:03 1.06 0.18 0.02 2/22/2017 0.07 0.69 23.99 0.28 1.35 1.22 90.50 0.01 22.91 1.74 59.85 0.18 0.15 1.18 90.90 0.09 0.29 2.12 0.08 1.36 0.14 0.01 0.01 23.09 0.42 0.08 1.47 1.20 91.34 0.09 24.68 2.18 82.36 0.01 0.09 0.09 0.39 3.11 0.07 1.32 0.06 0.01 25.82 0.20 0.44 87.57 1.71 0.09 84.60 0.10 2.65 0.01 0.09 0.21 0.45 1.43 0.12 0.05 3.56 0.01 0.49 88.69 0.08 89.98 0.08 0.08 3.21 0.01 0.49 1.22 0.15 0.08 0.05 0.01 89.10 0.99 0.15 0.08 0.06 0.01 0.06 90.53 2.05 0.15 0.05 0.07 90.70 2.41 0.14 90.62 0.09 0.06 2.38 0.12 90.29 0.06 2.33 0.19 91.39 0.07 2.49 91.49 0.06 2.23 91.53 0.06 91.52 0.06 91.52 91.37 1 OC7-RLB2-4W31c 4App Dil7-1 FP OC7-RLB2-3W520:20 2/20/2017 OC7-RLB2-3F1c 6App Dil7-1 FP 65.24 OC7-RLB2-2W71c App 21:21 Dil7-1 FP 2/20/2017 OC7-RLB2-2F1c 8App Dil7-1 FP 1.11 67.0521:51 2/20/2017 OC7-RLB2-1W91c App 22:22 Dil7-1 FP 2/20/2017 22.05 66.30 OC7-RLB2-1F1c App Dil7-1 FP 1.18 63.738:44 0C4-LFC-1W 2/21/2017 1c App 1.16Dil7-1 FP 1.389:14 2/21/2017 20.92 1.41 65.631c App Dil7-1 9:45 FP 2/21/2017 61.13 21.23 0.46 1.03 22.94 1.2110:15 2/21/2017 60.23 1.66 1.15 0.01 0.37 55.54 1.29 21.39 2.24 24.56 0.42 1.17 0.06 0.01 1.06 0.42 24.89 1.09 0.01 25.28 0.21 0.07 0.44 0.01 1.03 0.42 1.99 0.08 2.82 0.18 0.02 0.09 0.34 0.02 0.86 0.17 0.06 2.45 0.08 0.16 0.01 0.09 0.02 2.51 93.17 0.06 0.07 2.35 0.11 0.15 0.15 0.07 93.32 2.47 0.06 0.11 2.36 93.31 0.15 92.48 0.07 1.88 0.06 2.92 92.44 0.06 91.53 0.07 90.90 88.15 10 11 12 13 14 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 15ox Sample 1F=Top, Fresh 1F=Top, 2 OC7-RLB2-4F1c App Dil7-1 FP 20:50 2/20/2017 64.93 1.18 22.34 1.11 0.50 0.01 0.07 0.24 2.77 0.07 93.23

1W=Top, Weathered 1W=Top,

131

Flux Sample 7.06 0.850 6.000 7.06 0.850 5.999 7.067.06 0.850 0.850 6.000 6.000 7.06 0.850 6.000 7.0547.038 0.8517.058 0.853 6.000 7.057 0.850 6.002 0.850 6.000 6.001 7.0627.041 0.8507.064 0.852 6.000 7.065 0.849 6.000 7.061 0.849 6.000 7.063 0.850 6.000 0.850 5.999 7.059 6.002 7.065 0.850 0.849 6.000 7.062 6.000 0.850 6.000 7.0627.055 0.8507.061 0.851 6.000 7.056 0.850 6.001 0.850 6.000 6.000 7.059 0.850 6.000 7.059 0.850 6.000 7.0567.059 0.850 0.850 6.000 6.000 7.056 0.850 6.000 7.0597.059 0.850 0.850 6.000 6.000 dilution g g 50-50Flx Ig 9.34 9.16 9.26 7.396 7.456 8.537 10.674 mass% Zr 400 389 222 185 205 446 229 219.9 11.187 Y 26 29.3 362.1 6.703 24.1 28.8 211.3 8.964 26.9 240.5 15.198 15.9 27.7 274.2 8.408 27.9 24 Sr 33 36 27 28 110 19.3 27.834.9 226.8 33.649.1 10.718 189.6 43.4 9.349 213.5 8.519 34.9 33.2 292.8 8.497 Rb 171 190 191 199 60 115 170.4 22.6 22.7 229.7 11.734 113.9 19.5 36.8 416.4 8.342 Ce 144 129 93.3 118.5 14.3 21.2 396.7 6.556 134.8 139.9 Cr 85 89 75.4 142.7 115.1 24.7 25.3 387.4 6.568 96.4 95.5 96.4 32.8 42.5 Ni 11 29 13.5 103.7 Ba 298 ppm ppm ppm ppm ppm ppm ppm ppm Analysis Procedure Analysis Time / Date # Outcrop Sample Sample Outcrop OC4-LKC-1WOC4-LKC-1F1c App Dil7-1 FP 1c App Dil7-1 FP 15:03 3/17/2017 16:21 2/21/2017 128.2 296.5 11.9 31.2 57.8 88.6 181.9 158.2 161.2 141.4 34.4 44.1 30.6 25.4 321.0 17.534 256.1 12.312 0C4-LFC-1F0C4-LFC-2W0C4-LFC-2F1c App Dil7-1 FP 0C4-LFC-3W1c App Dil7-1 FP 0C4-LFC-3F10:46 2/21/2017 1c App Dil7-1 FP 13:49 2/21/2017 1c App Dil7-1 FP 292.714:19 2/21/2017 1c App Dil7-1 445.2 FP 14:50 OC4-LKC-2W 2/21/2017 79.1 584.3 41.515:20 2/21/2017 456.91c App 134.9Dil7-1 FP 41.7 116.5 512.4 29.2 113.2 16:52 2/21/2017 102.2 175.3 35.6 104.5 994.9 182.3 171.1 172.3 78.4 130.2 195.3 145.7 109.9 192.5 142.8 OC4-LKC-2FOC4-LKC-3WOC4-LKC-3F1c App Dil7-1 FP 1c OC4-LKC-4WApp Dil7-1 FP 19:55 2/21/2017 1c App Dil7-1 FP 20:25 2/21/2017 1c App Dil7-1 FP 300.68:57 2/22/2017 443.19:28 2/22/2017 45.3 34.2 289.5 457.5 119.5 107.8 35.7 142.1 30.8 113.9 112.4 145.2 122.7 186.8 182.6 104.3 117.1 65.6 170.6 214.4 86.4 16.5 43.7 189.5 24.2 26.2 9.895 215.6 10.776 OC4-LKC-4F1c App Dil7-1 FP 9:58 2/22/2017 453.9 32.7 108.3 163.9 OC4-LKC-5W1c App Dil7-1 FP 10:29 2/22/2017 217.7 OC4-LKC-5FOC4-LKC-6WOC4-LKC-6F1c App Dil7-1 FP 1c OC4-LKC-7WApp Dil7-1 FP 10:59 2/22/2017 1c App Dil7-1 FP 11:30 2/22/2017 1c App Dil7-1 FP 146.512:00 2/22/2017 330.312:31 2/22/2017 33.2 300.4 12.1 341.9 106.2 14.6 88.3 9.5 133.8 98.9 83.7 87.7 32.7 115.3 127.6 48.1 40.3 512.6 9.592 25.4 255.4 8.362 OC4-LKC-7F1c App Dil7-1 FP 13:01 2/22/2017 295.7 10.3 88.7 127.2 133.9 31.7 30.4 268.4 8.375 OC4-LKC-8W1c App Dil7-1 FP 13:32 2/22/2017 393.8 14.7 92.4 79.5 139.2 32.8 20.2 236.1 8.374 OC4-LKC-8F1c App Dil7-1 FP 14:03 2/22/2017 239.1 30.2 119.4 77.1 127.4 34 OC7-RLB2-3W5 OC7-RLB2-3F1c App Dil7-1 FP OC7-RLB2-2W1c App Dil7-1 FP 21:21 2/20/2017 1c App Dil7-1 FP 21:51 2/20/2017 397.922:22 2/20/2017 403.5 9.6 344.2 5.4 97.7 121.4 116.1 23.8 25.3 67 OC7-RLB2-2F8 OC7-RLB2-1W91c App Dil7-1 FP OC7-RLB2-1F1c App Dil7-1 FP 0C4-LFC-1W1c8:44 App 2/21/2017 Dil7-1 FP 9:14 2/21/2017 1c App Dil7-1 FP 197.99:45 2/21/2017 10:15 2/21/2017 12.3 270.3 369.5 80.5 11.2 43.4 143.5 101.9 102.9 113.5 122.5 25.1 97.9 36.2 37.8 45.8 558.8 8.96 16 10 11 12 13 14 17 18 19 20 21 22 23 24 25 26 27 28 29 30 15ox Sample 1F=Top, Fresh 1F=Top, 2 OC7-RLB2-4F1c App Dil7-1 FP 20:50 2/20/2017 441.3 6.8 74.6 83.4 126.9 33.6 25.1 350.9 6.647

1W=Top, Weathered 1W=Top, 1 OC7-RLB2-4W1c App Dil7-1 FP 20:20 2/20/2017 397.6 10.9 85.1 135.8 127.6

132

XRF Data Normalized to Total Mass 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 100 Normalized toMass Total Normalized SiO2 SiO2 TiO2 Al2O3 Fe2O3 MgO MnO CaO Na2O K2O P2O5 Total mass% mass% mass% mass% mass% mass% mass% mass% mass% mass% mass% FP Dil7-1 App 1c App Dil7-1 FP 69.64851c App Dil7-1 FP 1.26897 23.9635 71.84991c App Dil7-1 1.19174 FP 1.26876 0.54063 22.4176 71.05391c App 0.01523Dil7-1 1.10266 FP 1.48002 0.07187 0.39434 22.7523 68.91411c App 0.25959 0.01232Dil7-1 1.23246 FP 1.52902 2.97023 24.8061 71.0004 0.0793 0.44691c 0.06972 App Dil7-1 FP 1.30685 1.3971 0.01436 23.1403 66.7836 0.18861c 0.08359 0.45417App Dil7-1 1.14998 FP 1.81353 2.62325 0.17683 0.01449 0.47384 26.8314 66.2563 0.06322 1c 2.68462 0.10165App 0.01634Dil7-1 1.18972 FP 2.45862 0.07502 0.17518 0.08763 0.45884 27.3804 63.00611c 2.54224App 0.07681 0.01682Dil7-1 1.13636 FP 1.33182 0.06596 2.67644 0.09395 0.37842 28.6783 62.38271c App 0.16715 0.01375Dil7-1 2.25411 FP 0.0714 1.3403 2.57608 0.12211 0.98015 62.27151c 0.06883 App 27.5798 0.11881 0.01951Dil7-1 FP 1.33423 4.11174 0.16676 28.2206 61.1813 2.06921c 0.95111App 0.17016Dil7-1 FP 1.14363 3.3636 0.01839 3.31593 27.7564 62.3758 0.066 1c 0.23553 0.90032 0.07714 App Dil7-1 4.29164 FP 1.32342 0.13908 0.01805 1.14694 27.5685 61.2325 3.16736 0.10876 0.06353 1.17363 3.2838 0.07402 0.22648 0.22099 26.92131c 3.47236 0.92629App 0.19337Dil7-1 4.48104 FP 0.08409 0.01947 3.88502 1.18458 60.78591c 0.06601 0.11713 App 0.04642Dil7-1 FP 1.19562 0.36523 0.38647 33.3791 57.39041c 3.97247App 0.24305Dil7-1 1.92418 FP 1.15845 0.09901 4.13618 0.49218 30.9116 60.61711c 0.19488 App 0.01564Dil7-1 7.05944 FP 1.22452 0.05139 0.57213 32.9244 61.57161c App 0.06737 0.02128Dil7-1 1.90443 FP 1.1503 1.99041 0.06383 0.51078 61.26941c 0.09821 App 30.4191 0.01229Dil7-1 FP 1.19531 0.2175 0.07893 31.4035 62.1657 2.29061c 2.50366App 0.16349Dil7-1 2.22788 FP 1.11894 0.77465 0.10166 2.46257 0.62066 29.7684 62.78691c 0.10148 App 0.01717Dil7-1 1.84575 FP 1.24802 0.015 0.08754 0.76658 29.4365 70.67791c App 0.01348Dil7-1 1.83234 FP 0.07891 0.1055 2.1231 0.09168 0.75741 0.15004 72.0791c 2.97873App 23.9124 0.21539Dil7-1 FP 3.45535 0.09428 0.014 1.35176 2.41888 68.4118 0.09447 3.93121c 0.30566App 22.6271 Dil7-1 FP 1.47066 0.07607 0.08284 0.01368 1.17621 25.638 67.561 0.230421c 0.32119App Dil7-1 FP 3.53348 0.064 0.01418 1.34116 68.6908 1.4816 0.08489 1c 0.10522App 26.2221 Dil7-1 FP 1.28706 0.13462 0.45411 1.48545 25.0311 68.6711 0.0886 1.347351c 0.42191App Dil7-1 1.60501 0.0139FP 1.33081 0.06952 1.09536 0.01497 0.48511 25.2286 66.3709 0.05471 0.07421 1c 0.09072App 0.01431Dil7-1 1.43679 FP 1.28393 0.16195 0.16068 0.05026 0.49386 65.5019 2.24319 26.968 2.63423 0.16061 0.01442 1.31332 0.07003 0.06777 1.87071 2.60144 0.06993 28.2583 0.53105 0.15078 1.56942 0.0743 0.01311 2.54143 0.5308 0.04917 0.06228 0.12566 0.01357 2.71865 0.09522 0.06884 0.21013 2.44059 0.06676 FP Dil7-1 App 1c App Dil7-1 FP 60.7953 1.22392 32.7228 2.41748 0.49661 0.01627 0.04128 0.15178 2.03015 0.10442 FP Dil7-1 App 1c App Dil7-1 FP 70.0239 1.19032 23.6669 1.24291 0.49266 0.01524 0.06547 0.22003 3.02356 0.05903 0C4-LFC-1F 0C4-LFC-2W 0C4-LFC-2F 0C4-LFC-3W 0C4-LFC-3F OC4-LKC-1F OC4-LKC-2W OC4-LKC-2F OC4-LKC-3W OC4-LKC-3F OC4-LKC-4W OC4-LKC-4F OC4-LKC-5W OC4-LKC-5F OC4-LKC-6W OC4-LKC-6F OC4-LKC-7W OC4-LKC-7F OC4-LKC-8W OC4-LKC-8F 234 OC7-RLB2-4F 5 OC7-RLB2-3W 6 OC7-RLB2-3F 7 OC7-RLB2-2W 8 OC7-RLB2-2F 9 OC7-RLB2-1W OC7-RLB2-1F 0C4-LFC-1W 1 OC7-RLB2-4W 10 11 12 13 14 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 15ox OC4-LKC-1W

133

XRF Standards -2.14 -0.12 -1.46 3.45 0.37 88.725 0.03 -1.58 2.14 0.35 0.37 1.58 -0.48 6.05 -3.25 0.69 0.40 0.45 0.02 0.19 0.08 0.06 0.0033 0.100.19 0.20 0.01 0.02 0.13 0.01 0.03 1.08 0.04 0.0018 0.04 0.12 0.02 0.01 0.30 -1.9946.5 -2.21 0.838 -0.62 20.8 -3.11 9.356 -2.25 2.495 0.1425 -4.73 2.853 -2.15 0.415 3.35 0.356 87.1 SiO2 TiO2 Al2O3 Fe2O3 MgO MnO CaO Na2O K2O P2O5 Total 49.9049.92 2.73 2.69 13.50 13.79 12.30 12.34 7.23 7.26 0.167 0.1692 11.40 11.35 2.2246.69 2.35 0.52 0.84 0.50 20.87 0.27 0.27 100.24 9.41 100.64 2.54 0.14 2.89 0.31 3.38 0.37 87.43 50.3450.26 2.6950.40 2.6950.32 14.00 2.7249.52 13.92 2.70 12.4249.52 13.99 2.67 12.3549.57 7.31 13.92 2.70 12.4449.40 7.23 13.71 2.69 12.43 0.172 7.30 13.61 2.64 12.29 0.167 7.27 11.45 13.67 12.30 0.174 7.21 11.40 13.49 12.30 2.55 0.169 7.16 11.46 12.23 2.52 0.166 7.32 11.43 0.53 2.53 0.173 7.26 11.2746.46 0.52 2.55 0.166 11.25 0.2846.51 0.53 0.857 2.19 0.167 11.27 0.2846.56 0.52 101.73 0.839 2.17 20.87 11.22 0.2946.78 0.48 101.32 0.824 2.19 20.79 0.2846.89 0.48 9.38 101.84 0.852 2.13 20.62 0.2646.99 9.389 0.49 101.59 0.827 20.82 2.508 0.2646.82 0.49 9.39 0.832 2.533 99.76 20.99 0.2646.72 0.1394 9.417 0.841 99.62 20.98 2.543 0.144 0.26 9.454 2.847 0.815 2.551 99.93 21.05 0.1426 9.426 2.842 2.552 99.29 0.451 20.9 0.142 9.433 2.856 2.63 0.442 0.1451 3.359 2.566 2.9 9.43 0.423 2.926 3.367 0.145 0.365 0.1437 3.368 2.496 0.207 0.206 0.371 2.917 2.927 87.2 0.369 0.1418 3.389 3.38 87.2 0.191 0.264 2.911 87.1 0.368 3.399 0.373 3.384 0.194 87.7 0.372 0.368 87.4 3.389 87.9 87.8 0.384 87.4 mass% mass% mass% mass% mass% mass% mass% mass% mass% mass%% mass TRUE 47.64 0.86 21.00 9.71 2.60 0.15 2.95 TRUE Ave. Meas. StDev (wt. %) Ave. Meas. % Accuracy StDev (wt. %) date SBC-1SBC-1SBC-1App1c Dil7-1 FP SBC-1App1c Dil7-1 FP 11:47 2/21/2017 App1c Dil7-1 FP 17:53 2/21/2017 App1c Dil7-1 FP 19:49 2/20/2017 15:04 2/22/2017 SBC-1 C SBC-1 C SBC-1 C SBC-1 App1c Dil7-1 FP C SBC-1 App1c Dil7-1 FP 8:13 C SBC-1 2/21/2017 App1c Dil7-1 FP 12:17 2/21/2017 App1c Dil7-1 FP 18:23 2/21/2017 App1c Dil7-1 FP 18:18 2/20/2017 15:34 2/22/2017 BHVO-2 E BHVO-2 E BHVO-2 E BHVO-2 App1c Dil7-1 FP E BHVO-2 App1c Dil7-1 FP 13:18 2/21/2017 App1c Dil7-1 FP 19:24 2/21/2017 App1c Dil7-1 FP 17:17 2/20/2017 16:35 2/22/2017 BHVO-2 A BHVO-2 A BHVO-2 A BHVO-2 App1c Dil7-1 FP A BHVO-2 App1c Dil7-1 FP 11:16 2/21/2017 App1c Dil7-1 FP 17:22 2/21/2017 App1c Dil7-1 FP 19:19 2/20/2017 14:33 2/22/2017 Sample /Bead Sample BHVO-2 basalt Analy.Type SBC-1 Creek Brush shale AnalysesOhioUniversity at usingwere the theand Supermini200 FundamentalRigakuon conducted Parameters(FP) calibratedmethodagainst and Creek(Brush (basalt) Shale)SBC-1 BHVO-2 and standards. were duringanalysis. running usedstandards rock as USGS 12 Reference X-Rayfluorescence benchtopanalysisX-Rayfluroescence-using a Kansai,ofrocks K., 2008, spectrometer, Supermini. Journal,RigakuThe24(1)

134

Ig 0.1 0.3 0.6 -0.2 -1.9 -1.5 -2.0 -1.7 -0.1 12.71 12.59 12.61 12.73 12.41 12.09 11.96 12.03 12.45 mass % mass Zr 134 Y 36.5 Sr 389 26 172 429423 26 34 174 171 6 2 Rb 9.8 Cr Ce 109 108 147 178 -7.17 34.80 9.27 -16.51 -10.90 8.50 Ni 96 278 101 111 275 111 14 442 28 171 100 292 67 130 119 280 38 815884 109 9568 120 279 270 101 294 109 106 281 73 14 17 116 442 15 442 15 30 441 28 436 170 34 180 32 177 167 90 98 278 121 10 436 25 178 Ba 888735771 78700 74679 73 86737 75 107637 68 92 100685 70 194 94738 65 121 123 158 60 167 98 161 69 102 155 156 148 108 147 148 159 102 157 161 154 32 138 31 157 149 136 161 147 31 145 164 150 143 33 163 148 40 150 151 33 142 146 32 140 31 151 30 137 149 152 76.313.1 103.7 280.9 8.7 100.4 8.1 11.4 20.0 436.3 5.2 29.8 6.8 173.6 174.6 3.4 4.6 5.0 n.d. n.d. n.d. -41.3 -12.9 0.3 164.2 16.5 12.1 14.4 0.9 -7.38 ppm ppm ppm ppm ppm ppm ppm ppm 71.57 5.49 10.26 26.24 3.51 2.45 2.80 5.28 0.30 729.84 70.19 101.19 145.59 160.62 148.61 32.52 145.39 12.40 788.00

135

APPENDIX 6: CIA AND MAP CALCULATIONS Upper Freeport Raw CIA-K CIA-K Calculation From "Raw Data" CIA-K= (Al2O3/Al2O3+CaO+Na2O)*100 CIA-K (Weathered) CIA-K (Fresh) Depth (cm) 98.01 97.67 0 98.02 97.47 -25 97.45 96.07 -50 Upper Freeport Paleosol MAP MAP Calculation From CIA-K P= 14.265(CIA-K)-37.632 MAP (Weathered) MAP (Fresh) Depth (cm) 1360.44 1355.68 0 1360.61 1352.81 -25 1352.52 1332.75 -50

136

Lower Kittanning Paleosol CIA-K CIA-K Calculation From "Raw Data" CIA-K= (Al2O3/Al2O3+CaO+Na2O)*100 CIA-K (Weathered) CIA-K (Fresh) Depth (cm) 99.02 99.39 0 98.49 98.76 -25 98.73 98.95 -50 98.28 98.27 -75 98.61 98.53 -100 98.59 98.39 -125 98.60 98.53 -150 98.91 98.20 -175 Lower Kittanning Paleosol MAP MAP Calculation From CIA-K P= 14.265(CIA-K)-37.632 MAP (Weathered) MAP (Fresh) Depth (cm) 1374.85 1380.19 0 1367.33 1371.22 -25 1370.80 1373.91 -50 1364.33 1364.23 -75 1369.00 1367.93 -100 1368.80 1365.89 -125 1368.89 1367.97 -150 1373.37 1363.15 -175

137

Brookville Paleosol MAP MAP Calculation From CIA-K P= 14.265(CIA-K)-37.632 MAP (Weathered) MAP (Fresh) Depth (cm) 1375.12 1376.43 0 1373.12 1378.80 -25 1372.02 1372.72 -50 1371.86 1369.41 -75 Brookville Paleosol CIA-K CIA-K Calculation From "Raw Data" CIA-K= (Al2O3/Al2O3+CaO+Na2O)*100 CIA-K (Weathered) CIA-K (Fresh) Depth (cm) 99.04 99.13 0 98.90 99.29 -25 98.82 98.87 -50 98.81 98.64 -75

138

CIA-K and MAP Error Bar Calculations

Upper Freeport

Upper Freeport Paleosol Error Bar Calculations +Er=Molar proportion +Absolute Accuracy (wt%) CIA-K (+Er)= ((Al2O3+.103)/((Al2O3+.103)+(CaO+.035)+(Na2O+.15)*100 (W) F Depth (cm) 96.99 96.62 0 96.98 96.43 -25 96.41 95.03 -50

Upper Freeport Paleosol Error Bar Calculations -Er=Molar proportion -Absolute Accuracy (wt%) CIA-K (-Er)= ((Al2O3-.103)/((Al2O3-.103)+(CaO-.035)+(Na2O-.15)*100 (W) F Depth (cm) 99.05 98.75 0 99.08 98.54 -25 98.53 97.13 -50

How much CIA-K varies with the different errors (CIA-K+Er)-(CIA-K-Er)=Total Error Total Weathered Error Bar Total Error/2 Total Fresh Error Bar Total Error/2 2.07 1.03 2.13 1.07

2.10 1.05 2.11 1.05

2.12 1.06 2.10 1.05

2.09 1.05 2.11 1.06 The value above is the 'error bar' in the The value above is the 'error bar' in the line line graphs graphs

Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%)

(Al2O3+.106wt%)/Molecular Wt (CaO+.035wt%)/Molecular Wt (Na2O+.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.28231 0.27154 0.00360 0.00482 0.00517 0.00466 0.27782 0.27327 0.00256 0.00456 0.00607 0.00554 0.27143 0.26508 0.00180 0.00752 0.00831 0.00634

Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%)

(Al2O3-.106wt%)/Molecular Wt (CaO-.035wt%)/Molecular Wt (Na2O-.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.28023 0.26946 0.00235 0.00358 0.00033 -0.00018 0.27574 0.27119 0.00132 0.00332 0.00123 0.00070 0.26935 0.26300 0.00055 0.00627 0.00347 0.00150

Upper Freeport MAP Error Bar Calculation MAP(Error)= [(CIA-K Total Error)/100)]*MAP Total Weathered Error Error bar (one side)=Total Error/2 Total Fresh Error Error bar (one side)=Total Error/2 Depth (cm) 28.10 14.05 28.88 14.44 0 28.56 14.28 28.50 14.25 -25 28.67 14.34 27.97 13.99 -50 28.44 14.22 28.45 14.22 The highlighted value above is the average The highlighted value above is the average 'error' 'error' used in the MAP estimates of this used in the MAP estimates of this study and study and represents one side of the error represents one side of the error bars in the line bars in the line graphs. graphs.

139

Lower Kittanning

LKC Paleosol Error Bar Calculations +Er=Molar proportion +Absolute Accuracy (wt%) CIA-K (+Er)= ((Al2O3+.29)/((Al2O3+.29)+(CaO+.10)+(Na2O+.20)*100 (W) F Depth (cm) 98.10 98.49 0 97.53 97.86 -25 97.76 98.00 -50 97.29 97.27 -75 97.37 97.23 -100 97.44 97.27 -125 97.42 97.37 -150

97.81 97.16 -175

LKC Paleosol Error Bar Calculations -Er=Molar proportion -Absolute Accuracy (wt%) CIA-K (-Er)= ((Al2O3-.29)/((Al2O3-.29)+(CaO-.10)+(Na2O-.20)*100 (W) F Depth (cm) 99.96 100.32 0 99.47 99.69 -25 99.74 99.93 -50 99.29 99.30 -75 99.88 99.88 -100 99.78 99.55 -125 99.82 99.74 -150 100.05 99.27 -175

How much CIA-K varies with the different errors (CIA-K+Er)-(CIA-K-Er)=Total Error Total Weathered Error Bar Total Error/2 Total Fresh Error Bar Total Error/2

1.85 0.93 1.83 0.92

1.94 0.97 1.83 0.92

1.98 0.99 1.93 0.96 2.00 1.00 2.02 1.01 2.51 1.26 2.65 1.33 2.34 1.17 2.28 1.14 2.40 1.20 2.38 1.19 2.24 1.12 2.11 1.05 2.16 1.08 2.13 1.06 The value above is the 'error bar' in the The value above is the 'error bar' in the line line graphs graphs

Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%)

(Al2O3+.106wt%)/Molecular Wt (CaO+.035wt%)/Molecular Wt (Na2O+.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.321977549 0.328414333 0.001360253 0.001540436 0.00486892 0.003507178 0.304213682 0.323954946 0.001762354 0.002031535 0.005929403 0.005057997 0.299382639 0.309038027 0.002031196 0.00218516 0.004840903 0.004122321 0.293001313 0.289746037 0.002258917 0.001980597 0.005895337 0.006137796 0.235566989 0.222961363 0.001765368 0.002500299 0.0045922 0.003849687 0.252491403 0.258219931 0.001599714 0.002241843 0.005033033 0.005012542 0.24653895 0.248475637 0.001520308 0.001871035 0.005011454 0.004852879 0.26553513 0.27819042 0.001500922 0.002321965 0.004447581 0.005810445

Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%)

(Al2O3-.106wt%)/Molecular Wt (CaO-.035wt%)/Molecular Wt (Na2O-.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.319898303 0.326335087 0.000112036 0.000292219 2.86488E-05 -0.001333093 0.302134435 0.321875699 0.000514137 0.000783318 0.001089132 0.000217726 0.297303392 0.30695878 0.00078298 0.000936943 6.32089E-07 -0.00071795 0.290922066 0.28766679 0.001010701 0.00073238 0.001055066 0.001297525 0.233487742 0.220882117 0.000517151 0.001252082 -0.000248071 -0.000990584 0.250412157 0.256140684 0.000351497 0.000993627 0.000192762 0.000172271 0.244459703 0.24639639 0.000272091 0.000622818 0.000171183 1.2608E-05 0.263455884 0.276111174 0.000252706 0.001073748 -0.00039269 0.000970174 140

Lower Kittanning MAP Error Bar Calculation MAP(Error)= [(CIA-K Total Error)/100)]*MAP Total Weathered Error Error bar (one side)=Total Error/2 Total Fresh Error Error bar (one side)=Total Error/2 Depth (cm) 25.49 12.74 25.31 12.65 0 26.50 13.25 25.11 12.56 -25 27.16 13.58 26.50 13.25 -50 27.32 13.66 27.62 13.81 -75 34.40 17.20 36.26 18.13 -100 32.07 16.03 31.14 15.57 -125 32.85 16.43 32.52 16.26 -150 30.82 15.41 28.70 14.35 -175 29.58 14.79 29.15 14.57 The highlighted value above is the average The highlighted value above is the average 'error' 'error' used in the MAP estimates of this used in the MAP estimates of this study and study and represents one side of the error represents one side of the error bars in the line bars in the line graphs. graphs.

141

Brookville

Brookville Paleosol Error Bar Calculations +Er=Molar proportion +Absolute Accuracy (wt%) CIA-K (+Er)= ((Al2O3+.106)/((Al2O3+.106)+(CaO+.035)+(Na2O+.15)*100 (W) F Depth (cm) 97.27 97.42 0 96.95 97.50 -25 96.72 96.81 -50 96.78 96.52 -75

Brookville Paleosol Error Bar Calculations -Er=Molar proportion -Absolute Accuracy (wt%) CIA-K (-Er)= ((Al2O3-.29)/((Al2O3-.29)+(CaO-.10)+(Na2O-.20)*100 (W) F Depth (cm) 99.50 99.61 0 99.35 100.11 -25 99.36 99.42 -50 99.28 98.97 -75

How much CIA-K varies with the different errors CIA-K(+Er) - CIA-K(-Er)= Total Error Total Weathered Error Bar Total Error/2 Total Fresh Error Bar Total Error/2 2.23 1.11 2.19 1.09 2.39 1.20 2.61 1.30 2.64 1.32 2.61 1.30 2.50 1.25 2.46 1.23 2.44 1.22 2.46 1.23 The value above is the 'error bar' in the The value above is the 'error bar' in the line line graphs graphs

Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%) Molar Proportion + Absolute Accuracy (wt.%)

(Al2O3+.106wt%)/Molecular Wt (CaO+.035wt%)/Molecular Wt (Na2O+.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.264196181 0.269580058 0.002299461 0.002801465 0.005116978 0.00433698 0.244331706 0.227994409 0.002436634 0.002186664 0.005246498 0.003659403 0.220906319 0.224188534 0.002038115 0.002114709 0.005463043 0.00527317 0.233158822 0.236067847 0.001791601 0.001905652 0.005970186 0.006608365

Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%) Molar Proportion - Absolute Accuracy (wt.%)

(Al2O3-.106wt%)/Molecular Wt (CaO-.035wt%)/Molecular Wt (Na2O-.15wt%)/Molecular Wt

Al2O3(W) Al2O3(F) CaO(W) CaO(F) Na2O(W) Na2O(F) 0.262116934 0.267500811 0.001051244 0.001553248 0.000276707 -0.000503291 0.242252459 0.225915162 0.001188418 0.000938447 0.000406227 -0.001180868 0.218827073 0.222109287 0.000789898 0.000866492 0.000622772 0.000432899 0.231079575 0.2339886 0.000543384 0.000657435 0.001129915 0.001768094

Upper Freeport MAP Error Bar Calculation MAP(Error)= [(CIA-K Total Error)/100)]*MAP Total Weathered Error Error bar (one side)=Total Error/2 Total Fresh Error Error bar (one side)=Total Error/2 Depth (cm) 30.62 15.31 30.13 15.06 0 32.88 16.44 35.95 17.98 -25 36.26 18.13 35.81 17.90 -50 34.34 17.17 33.62 16.81 -75 33.52 16.76 33.88 16.94 The highlighted value above is the average The highlighted value above is the average 'error' 'error' used in the MAP estimates of this used in the MAP estimates of this study and study and represents one side of the error represents one side of the error bars in the line bars in the line graphs. graphs.

! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! ! !

Thesis and Dissertation Services ! !