<<

Clay Mineralogy, Provenance, and Sequence Stratigraphy of Upper in Eastern

Ohio

A thesis presented to

the faculty of

the College of Arts and Sciences of University

In partial fulfillment

of the requirements for the degree

Master of Science

Devin R. Fitzgerald

April 2016

© 2016 Devin R. Fitzgerald. All Rights Reserved. 2

This thesis titled

Clay Mineralogy, Provenance, and Sequence Stratigraphy of Upper Ordovician Shales in Eastern

Ohio

by

DEVIN R. FITZGERALD

has been approved for

the Department of Geological Sciences and

The College of Arts and Sciences by

Gregory Nadon

Associate Professor of Geological Sciences

Robert Frank

Dean, College of Arts and Sciences 3

ABSTRACT

FITZGERALD, DEVIN R., M.S., April 2016, Geological Sciences

Clay Mineralogy, Provenance, and Sequence Stratigraphy of Upper Ordovician Shales in Eastern

Ohio.

Director of Thesis: Gregory Nadon

A combination of x-ray diffraction analyses of core data and spectral gamma ray logs were used to interpret the largely succession of the Late Ordovician between the top of the Trenton and the Queenston Shale in the subsurface of east-central Ohio. The four county study area is within the back-bulge region of the foreland basin associated with the

Taconic Orogeny. The XRD data from the basal portion of the section reveal an increase upward in chlorite and along with a decrease in carbonate, which is consistent with an increase in detrital rather than authigenic clays. Detrital chlorite is a common clay in sediments shed from mountain belts and the appearance of the clay allows constraints to be placed on the transition from under to overfilled foreland basin. Correlation of six 4th-order sequences from

300 wells allowed construction of isopach maps throughout the 450 m interval. The isopach maps show that subsidence across the area was consistent and augmented by some combination of compaction over pre-existing structural and depositional features. 4

ACKNOWLEDGMENTS

The author wishes to recognize first and foremost, Eastern Mountain Fuel, the company that has employed and paid for this research project and a full time opportunity on completion of the research. Eastern Mountain Fuel’s owner/geologist Bob Thomas has provided me with guidance, opportunity, and encouragement. I appreciate his never ending enthusiasm for the industry and academic side of Geology. I also want to thank geologist Tom Thomas (EMF) for his support and help when problems arise with my GeoGraphix Software, and for being extremely helpful in creating the files to be easily used and analyzed within the data section. Eastern

Mountain Fuel has employed me since 2011, and I am extremely grateful for all the opportunities they have given me. I also want to acknowledge Artex Energy Group for helping to provide all data necessary in this research project including funding. I look forward to working as a team over many years to come in the Appalachian Basin.

In addition to my employers this thesis would not have been possible without the assistance of

Dr. Greg Nadon, my advisor on the research project. I thank you for being particularly patient with me over the last two years in the finishing of this project. Your knowledge of sedimentology, tectonics, and sequence stratigraphy has been wonderful in helping me develop my skills. A big thank you also goes out to my committee members Dr. Kidder and Dr. Nance for their advice on the research. Another thank you to Ohio University professor Dr. Gierlowski-

Kordesch for making me read and write so many papers that advanced my knowledge of broad, complex sedimentary processes and environments. Lastly, I would like to acknowledge my friends and family for being so persistent with me during my time at Ohio University and my entire academic career. Without your love and continuing support I would not be in the position I am today. 5

TABLE OF CONTENTS

Page

Abstract………………………………………………………………………………………………………………………………………3

Acknowledgments……………………………………………………………………………………………………………………...4

List of Tables……………………………………………………………………………………………………………………………….7

List of Figures………………………………………………………………………………………………………………………………8

Chapter 1: Introduction……………………………………………………………………………………………...... 10

Chapter 2: Previous Studies………………………………………………………………………………………………………13

2.1 Introduction……………………………………………………………………………………………………………13

2.2 Upper Ordovician Stratigraphy of East-Central Ohio……………………………………………….13

2.2.1 Trenton/Lexington Limestone………………………………………………………………...13

2.2.2 Point Pleasant Shale………………………………………………………………………………..17

2.2.3 ………………………………………………………………………………………………17

2.2.4 Kope Formation………………………………………………………………………………………19

2.2.5 Cincinnati Group……………………………………………………………………………………..19

2.2.6 Queenston Shale…………………………………………………………………………………….20

2.3 Sequence Stratigraphy……………………………………………………………………………………………20

2.4 Tectonics………………………………………………………………………………………………………………..23

2.5 Clay Mineralogy………………………………………………………………………………………………………25

Chapter 3: Methodology……………………………………………………………………………………………………………30

3.1 Introduction……………………………………………………………………………………………………………30

3.2 Well Logs………………………………………………………………………………………………………………..30

3.3 XRD Data………………………………………………………………………………………………………………..33 6

Page

3.4 Well Log Evaluation of Mineralogy………………………………………………………………………….34

Chapter 4: Results…………………………………………………………………………………………………………………….41

4.1 Introduction……………………………………………………………………………………………………………41

4.2 Correlation of Lithofacies……………………………………………………………………………………….41

4.3 Isopach Maps………………………………………………………………………………………………………….46

4.4 Sequence Stratigraphy……………………………………………………………………………………………57

4.5 XRD Mineral Abundance Data…………………………………………………………………………………60

4.6 Well Log Analysis of Clay Mineralogy………………………………………………………………………64

Chapter 5: Discussion………………………………………………………………………………………………………………..70

5.1 Introduction……………………………………………………………………………………………………………70

5.2 The Creation of Infilling of Accommodation……………………………………………………………70

5.3 Provenance…………………………………………………………………………………………………………….70

Chapter 6: Conclusions……………………………………………………………………………………………………………..76

References ………………………………………………………………………………………………………………………...... 77

Appendix A: XRD Data……………………………………………………………………………………………………………….82

Appendix B: Wells………………………………..……………………………………………………………………………………87

7

LIST OF TABLES

Page

Table 3.1: Summary of Log Functions………………………………………………………………………………………..40

Table 4.1: Average Thickness…………………………………………………………………………………………………….45

Table 4.2: XRD LePage……………………………………………………………………………………………………………….63

8

LIST OF FIGURES

Page

Figure 1.1: Map of Study Area…………………………………………………………………………………………………..12

Figure 2.1: Stratigraphic Equivalents in the Appalachian Basin………………………………………………….14

Figure 2.2: Depositional Setting…………………………………………………………………………………………………15

Figure 2.3: LePage #1 Cored Interval…………………………………………………………………………………………18

Figure 2.4: Sequence Stratigraphic Diagram………………………………………………………………………………21

Figure 2.5: Contacts of the clastic wedge of the Upper Ordovician…………………….……………………..22

Figure 2.6: Foreland Basin Evolution…………………………………………………………………………………………24

Figure 2.7: Dynamic Subsidence……………………………………………………………………………………………....26

Figure 2.8: Clay Chart………………………………………………………………………………………………….28

Figure 3.1: Map of Data Points………………………………………………………………………………………………….31

Figure 3.2: Example of XRD Data……………………………………………………………………………………………….35

Figure 3.3: Quirein et al. (1982) Clay Mineralogy Chart……………………………………………………………..36

Figure 3.4: Corrected Gamma-Ray…………………………………………………………………………………………….37

Figure 3.5: Study Outline…………………………………………………………………………………………………………..39

Figure 4.1: Utica/Point Pleasant Lithostratigraphy………………………………………………………………….…43

Figure 4.2: Average Percent Minerals……………………………………………………………………………………….44

Figure 4.3: Point Pleasant Isopach…………………………………………………………………………………………….47

Figure 4.4: Utica Isopach…………………………………………………………………………………………………………..48

Figure 4.5: Mud A Isopach…………………………………………………………………………………………………………49

Figure 4.6: Mud B Isopach…………………………………………………………………………………………………………50

9

Page

Figure 4.7: Mud C Isopach………………………………………………………………………………………………………...51

Figure 4.8: Mud D Isopach…………………………………………………………………………………………………………52

Figure 4.9: Queenston Isopach………………………………………………………………………………………………….53

Figure 4.10: Total Sequence Isopach…………………………………………………………………………………………54

Figure 4.11: Thickness Trends Summary …………………………………………………………………………………..55

Figure 4.12: Appalachian structural trends………………………………………………………………………………..56

Figure 4.13: Sequence Stratigraphy Concepts……………………………………………………………………………58

Figure 4.14: Upper Ordovician Sequence Stratigraphy………………………………………………………………61

Figure 4.15: Sequence Stratigraphy Cored Interval……………………………………………………………………62

Figure 4.16: Spectral Gamma-Ray Plots…………………………………………………………………………………….65

Figure 4.17: vs. Gamma-ray…….…………………………………………………………………………………………66

Figure 4.18: Clay minerals vs. Quartz……………………………………………………………………………………….68

Figure 4.19: Clay Minerals Plot…...... 69

Figure 5.1: Depositional Model…………………………………………………………………………………………………74

Figure 5.2: Global Sea Level………………………………………………………………………………………………………75

10

CHAPTER 1: INTRODUCTION

The sedimentary rock record is dominated by mudstones (Blatt, 1982). These fine- grained sedimentary rocks can serve as sources, reservoirs, and seals of hydrocarbons; influence the flow of groundwater; and can be rich in metals (Bohacs and Lazar, 2010). Investigation of

“shales” or “mudrocks” can provide information about provenance, climate, and sea level rise and fall through geological time (Bohacs and Lazar, 2010). The recent drilling for oil and gas resources in Ordovician (Utica/Point Pleasant) shales of the Appalachian basin has resulted in large amounts of new geophysical well log and core data of the predominantly shale section between the top of the Upper Ordovician Trenton Limestone and the Clinton Sandstone

Formation in eastern-central Ohio. Deposition of the Upper Ordovician siliciclastics was coeval with accretion of the terranes that produced the Taconic Orogeny (Pope and Read, 1997), and the shift in deposition style between the top of the Trenton Limestone to the top of the

Queenston Shale; representing a change from a clean carbonate system to a deltaic system with sediment infilling the foreland basin and onlapping onto the distal craton.

The accommodation that preserved the shales and mudstones above the Trenton

Limestone was a result of a combination of eustasy and subsidence associated with the Taconic

Orogeny (McLaughlin and Brett, 2007). At some point the foredeep that formed in response to thrust loading along the eastern margin of Laurentia was overfilled and sediments from the orogeny were dispersed in the back-bulge zone. A shallow, narrow foredeep would fill rapidly, whereas a broad deep moat would require more time to infill. The depth and width of a foreland basin is a function of the rigidity of the underlying lithosphere and the size of the load imposed. Previous work on the in southeastern Ohio (Muslim, 2014) showed the forebulge associated with the Acadian Orogeny was located in and 11

Pennsylvania. The size of the imposed load was smaller during the earlier Taconic Orogeny and the rigidity of the lithosphere was likely lower (DeCelles and Giles, 1996), therefore the Taconic foredeep was likely both narrow and shallow. Nevertheless, approximately 450 m of primarily mudstones were deposited in east-central Ohio coeval with the Taconic Orogeny.

The hypothesis that the Taconic foredeep was narrow and shallow can be tested using isopach maps and clay mineralogy. The inboard margin of a foredeep, which is marked by the peripheral bulge, migrates cratonward as the thrust sheets that form the orogeny advance

(DeCelles and Giles, 1996). Isopach maps on the foredeep side of the peripheral bulge should show an increase in thickness toward the axis of the foredeep, whereas those within the backbulge should not (Rodgers, 1971). The mineralogy of mudstones can indicate provenance provided the clays are detrital and not authigenic (Weaver, 1989). The abrupt increase in the amount of detrital chlorite, which is linked to erosion in orogenic belts (Weaver, 1989), would identify the switch in provenance from craton to orogeny.

This study tests three hypotheses. First, that clay mineralogy can be inferred from well logs. Second, that the detrital clay mineralogy shows a change in provenance from cratonic to

Appalachian within the mudstones. Third, that the accommodation for deposition was a result of dynamic subsidence. The first hypothesis requires access to core analyses to calibrate to well log data and these are available. The second requires determining the location in the section where detrital clays, such as chlorite appear. The third hypothesis can be tested using a combination of isopach maps and sequence stratigraphy determined from the well logs.

12

Figure 1.1: The study area in east-central Ohio includes portions of Guernsey, Morgan, Muskingum and Noble counties.

13

CHAPTER 2: PREVIOUS STUDIES

2.1 Introduction

The stratigraphic succession of the Upper Ordovician (458 Ma-444 Ma; Pope and Read,

1997) in the subsurface of eastern Ohio is a mixed carbonate-clastic system that is divided into six formations (Figure 2.1). The basal carbonate rich interval known as the Point Pleasant

Formation forms a transition from the of the Trenton and Lexington into a widespread, dark gray shale locally termed the Utica (Figure 2.2). The Utica is overlain in ascending order by mudstones and subordinate carbonates of the Kope Formation, the

Cincinnati Group, and the Queenston Shale (Smosna and Patchen, 1991; McClain, 2013; Patchen et al., 2015).

2.2 Upper Ordovician Stratigraphy of East-Central Ohio

The Upper Ordovician formations researched here have been extensively studied in surface exposure in areas of the Cincinnati Arch, Nashville Dome, and in Virginia. However, few published subsurface studies of the Late Ordovician strata include the east-central Ohio region.

Descriptions of each major lithologic units are provided below.

2.2.1 Trenton/Lexington Limestone

The Trenton Limestone (Vanuxem, 1838) is Upper Ordovician in age and consists of whole or fragmented , mainly crinoids, brachiopods, and bryozoans, in a fine, dark-gray to light brown limestone matrix with little siliciclastic mud content (Wickstrom et al., 1992). The

Trenton limestone lies above the in Ohio (Patchen et al., 2006). The Trenton

14

Figure 2.1: Modified from stratigraphic column of Appalachian Basin (Patchen et al., 2015).

15

Figure 2.2: Modified geological setting (Ettensohn, 2010) of the Appalachian Basin during the middle/upper Ordovician. Red Box is highlighting the study area.

16

Limestone lies above the Black River Group in Ohio (Patchen et al., 2006). This limestone is considered part of the Black River Group in Ohio (Patchen et al., 2006). The Trenton ranges in thickness from slightly less than 12 m in parts of western-central Ohio up to 90 m in northwestern Ohio (Wickstrom et al., 1992). The Trenton platform extends from southeast

Michigan and to (Patchen et al., 2006). The Trenton paleotopography was directly affected by the amount of erosion on the top of the underlying Beekmantown/Knox

Dolomite (Patchen et al., 2006). The Trenton structural highs correspond to remnants of the

Knox that were preserved below the Knox Unconformity.

The stratigraphically equivalent Lexington Limestone (Orton, 1890) crops out in

Kentucky and is present in the subsurface of Ohio as a series of interbedded limestone and shale members. The terms Trenton and Lexington are often used in the literature to describe the same strata. The limestones located in northern and eastern Ohio are termed the Trenton whereas limestone formations located in central and southern Ohio are often termed the

Lexington Limestone. The depositional environment at the time was dominated by abundant reef systems (Figure 2.2) with an elongated, slightly deeper sedimentary trough, termed the

Sebree Trough (Ettensohn, 2010), between the Trenton and Lexington Platforms. Carbonate deposition was able to keep pace with marine transgressions resulting in a shallow sea covering present- day Ohio. The two carbonate platforms restricted water movement within Ohio and a sub-basin with grey/black carbonaceous shale was deposited (Patchen et al., 2006). This limestone formation shows evidence of bioturbation throughout (Patchen et al., 2015). The carbonates were deposited in an environment where light, aeration, and nutrients were abundant and marine fauna thrived (Patchen et al., 2015).

17

2.2.2 Point Pleasant Shale

Orton (1890) was the first to give the name “Point Pleasant” to the strata exposed at

Point Pleasant, Clermont County, Ohio (Patchen et al., 2006). The Point Pleasant, which is equivalent to the Clays Ferry Formation of and the Dolgeville Formation of New York

(Patchen et al., 2006). Trenton and Lexington limestones laterally interfinger with the Utica shale and are overlain by the Point Pleasant Formation (Patchen et al., 2006). It is described as dark gray/black organic rich marl that is thinly interbedded and laminated with fossiliferous mudstones. The upper portion of both the Trenton and Lexington limestones laterally interfinger with the Utica shale and are overlain by the Point Pleasant Formation (Patchen et al.,

2006). The distribution of the shales of the Point Pleasant Formation is controlled by topography on top of the Trenton and Lexington Limestone. The Point Pleasant is approximately 30 m thick in eastern Ohio and composed of organic-rich, black calcareous shales alternating from 30% to 50% dark argillaceous limestones (Brett et al., 2004) in an overall fining upward sequence (Figure 2.3). The limestones include tempestite beds composed of crinoids and bryozoan fossils (Huck, 2013). The top of the Point Pleasant is placed above a regionally extensive limestone that is interpreted as the result of a laterally extensive regression (Patchen et al., 2015).

2.2.3 Utica Shale

The name Utica Shale was first used by Emmons (1842) to refer to the black shale in the

Mohawk Valley, New York that underlies the Lorraine Shales and overlies the Trenton

Limestone. The Utica shale occurs throughout much of the central Appalachian basin, extending across the Appalachian Plateau from New York and to southern and covering 18

Figure 2.3: Stratigraphic subdivision of the gamma-ray log of the LePage #1 well in Muskingum County, showing the cored interval from the top of the Utica Shale to 20 feet into the Trenton. The core was sampled for analysis every 5 feet.

19 approximately 28,000 square miles. The Utica can be traced into the fold-and-thrust belt of the

Appalachians where it is partly equivalent to the Antes Shale. In eastern Ohio the Utica Shale is a massive, fossiliferous, organic-rich, black to gray shale that ranges from 30 to 70 m thick and interfingers and overlies the Point Pleasant Formation (Ettensohn, 2010). The preservation of the organic material within the Utica Shale is interpreted to be a result of restricted circulation and low energy conditions stemming from a major transgression across the eastern United

States (Bergstrom and Mitchell, 1992; Ryder, 2008; Wickstrom, 1992).

2.2.4 Kope Formation

The Kope Formation is equivalent to the lower portion of the Calloway Creek Limestone of Kentucky, the of New York, and part of the Reedsville Shale of and West Virginia (Patchen et al., 2015). The Kope Formation in southeastern Ohio is 200 m in thickness (Patchen et al., 2006) and is composed of interbedded shale (70%), limestone, and siltstone. In east-central Ohio the unit is recognized by a very stable gamma ray signal above the Utica Shale and is approximately 75 m thick.

2.2.5 Cincinnati Group

The Cincinnati Group is a succession of gray to black, silty shales that is more than 230 m thick in the northwestern part of the basin (Wickstrom et al., 1992). The Cincinnati Group in

Ohio is equivalent to part of the Reedsville Shale in West Virginia and the Martinsburg

Formation of central Pennsylvania (Wickstrom et al., 1992). The Reedsville records a siliciclastic source area on the northeast that grades southward and westward through alluvial and flood plain environments into shallow marine limestones and shales (Fisher, 1977). 20

2.2.6 Queenston Shale

The uppermost Ordovician strata in the study area comprise the Queenston Shale, which grades eastward into continental red beds of the (Wickstrom et al.,

1992). The redbeds of the Queenston Shale in the subsurface of eastern Ohio can range up to

100 m in thickness (Cressman, 1973). The Queenston lies between the silty gray shales of the

Cincinnati Group and sandstones of the Medina Group that is Silurian in age. The contact between the Cincinnati Group and the Queenston Shale is based on a color change and therefore is difficult to determine without core or drill cuttings.

2.3 Sequence Stratigraphy

The Upper Ordovician section in the vicinity of the Cincinnati Arch, Nashville Dome, and

Lexington Kentucky has been subdivided into three 2nd-order and eight 3rd-order sequences that were the result of the combination of tectonism, terrigenous sediment input, and glacial eustasy

(Holland and Patzkowsky, 1996, 1997; Pope and Read 1997, 1998). The distance from the outcrop studies and the lack of core data available within the study interval prevents a precise correlation of the 3rd order sequences. The intervals mapped as sequences in this study were chosen based on gamma-ray deflections that could be correlated across the map area (Figure

2.4). The deflections are increases in total gamma-ray values that occur at the top of the overall decreasing upward trends and are interpreted as coarsening and shallowing upward of the section (Figure 2.5).

21

Figure 2.4: Sequence stratigraphic names from past studies, modified after (Holland and Patzkowsky, 1996). The units within the Cincinnati Group are informal and based on gamma-ray correlation.

22

Figure 2.5: Contacts of the clastic wedge of the Upper Ordovician based on gamma ray reflections on carbonate zones above the Utica Shale to the top of the Queenston. The labels are formation tops based on gamma-ray correlation.

23

2.4 Tectonics

During the Late Ordovician the paleolatitude of the North American craton was between

20 and 25 °S and rotated approximately 90° clockwise relative to its present position with a passive margin to the south (Brett et al., 2004). Beginning in the Middle Ordovician collision with a along the southern margin resulted in two main pulses, or tectophases

(Ettensohn, 2004) of thrust fault movement that together are termed the Taconic Orogeny. The tectonic activity of the earlier Blountian tectophase (Whiterockian-Mohawkian) was mostly focused in vicinity of present-day Georgia and Virginia (Bradley, 1989). During the subsequent

Taconic tectophase (Chatfieldian-Cincinnatian) activity shifted toward present day New York state (Ettensohn, 2010). The result of these events was the formation of foreland basins. Most authors suggest the second tectophase occurred during the Edenian (Kope Formation)

(Ettensohn, 2004), whereas Pope and Read (1997) and Holland and Patzkowsky (1996) infer a later impact.

The result of the thrust fault movements was the development of the first stage of the

Appalachian foreland basin. A foreland basin is an elongate trough that forms on continental crust between an orogenic belt and an adjacent craton due to thrust loading, and is divided into four depozones; the wedge top, foredeep, forebulge, and back-bulge (Figure 2.6)(DeCelles and

Giles, 1996). The width and depth or elevation of the different zones is a function of the width and thickness of the thrust loads and the strength of the underlying lithosphere. If the rate of formation of accommodation in the foredeep exceeds the sedimentation rate the basin is termed underfilled and no sediment from the orogeny reaches the backbulge region (DeCelles and Giles, 1996). Only when the foredeep is overfilled will sediment from the orogeny overtop the forebulge. The total accommodation available for sediment and water in the back-bulge is 24

Figure 2.6: The foreland through the Upper Ordovician modified after DeCelles and Giles (1996) and Castle (1998).

25 limited by minor flexural responses cratonward of the forebulge (DeCelles and Giles, 1996).

Back-bulge basins most likely develop where a strong viscous coupling exists between the base of the continental plate and downward circulating mantle-wedge material that becomes entrained by the subducting oceanic plate (i.e., dynamic subsidence) (DeCelles, 2011).

The size of the Taconic foredeep is not known but some constraints can be placed on the system through modeling. Muslim (2014) estimated the location of the forebulge of the

Devonian Acadian Orogeny using published estimates of the axis of the foredeep and modern lithosphere rigidity. He found that the maximum westward location of a forebulge would lie within central West Virginia and Pennsylvania. The smaller thrust loads during the Taconic and, presumably, warmer lithosphere means that the location of the earlier Taconic forebulge must lie to the east of that calculated for the Devonian. Muslim’s (2014) study estimates that the distance from the foredeep to the backbulge is approximately 490 km in the Appalachian Basin

(Figure 2.7).

2.5 Clay Mineralogy

The minerals of the common sedimentary rocks can be divided into four major groups:

(1) minerals that survive weathering and transport (detrital minerals), (2) new minerals formed during weathering and transportation (secondary minerals), (3) minerals that form directly from solutions chemically or biochemically (precipitated minerals), and (4) authigenic minerals that form in sediments during and after deposition (Quirein et al., 1982). Only the detrital minerals are used for provenance analysis. Mudstones are composed primarily of detrital clay minerals with lesser amounts of quartz and feldspar. In marine sediments, clays are primarily detrital

26

Figure 2.7: Foreland basin system during the middle Ordovician. The study area is shown craton- ward. The arrows represent the subsidence taking place (dynamic and thrust). The distance from the foredeep to the backbulge is approximately 490 km (after DeCelles and Giles 1996 and Muslim, 2014).

27

(90%; Velde, 1995) and the abundance of the difference types provides abiotic proxy data to determine climate changes in the source area, changes in weathering intensity, and changes in the terrigenous sediment input (Gingele et al., 2004).

The most common clay minerals in mudstones are illite, smectite, chlorite, and

(Figure- 2.8). The type of present is determined using X-Ray diffraction (XRD) because the lattice spacing of the different minerals varies. Illite is the term applied to the most common clay. Illite belongs to the group and is a primary product of the weathering of silicates and degradation of (Weaver, 1989). Detrital illite primarily forms by hydrolysis in semi-arid climates (Chamley, 1989). Smecite is the term applied to a clay mineral that has an angle reflection >17°, in regards to the crystal lattice spacing. Smectites are usually a strong indicator of rock within the source area (Weaver, 1989) and can form in-situ under hydrothermal activity. In addition, smectite often exists as mixed-layered clay in the study area.

During diagenesis smectite will be go through a process called illitization and transform into illite at depths of 10,000-15,000 feet (Weaver, 1989). The is an Fe-rich clay with a 14° angle reflection under XRD. Chlorite is a widespread primary constituent of low-grade metamorphic, magmatic, terrigenous sedimentary rocks, and a secondary weathering product of other clay minerals such as illite (Weaver, 1989).

Most of the chlorite in shales is detrital in origin, particularly in the facies

(Weaver, 1989; Velde, 1995). Chlorite is relatively stable and usually is transported to sea by river and deltaic systems. Authigenic chlorite is relatively unstable and forms in shallow soils and sandstone formations. Kaolinite is identified by a 7° angle reflection that disappears above

28

Figure 2.8: Diagram of clay minerals that are common in the study area, indicators of how they occur, and a range of their occurrence in the Utica/Point Pleasant Interval (Weaver, 1989).

29

500 °C. Kaolinite and pure smectite in the study area is alternated by diagenesis during burial, the most dominant clay minerals in the Point Pleasant Formation, Utica Shale, Cincinnati Group, and Queenstone Shale are illite and chlorite clay minerals.

30

CHAPTER 3: METHODOLOGY

3.1 Introduction

This study will combine geophysical well log data with XRD data from cored intervals to interpret the depositional environment of the stratigraphic section lying above the Trenton

Limestone. The bulk of the data available are within the Point Pleasant Shale from wells in

Guernsey, Muskingum, Noble, and Morgan Counties (Figure 3.1). This data will be used to construct a model that will be analyzed and tested using clay mineralogy, statistical methods, and mapping techniques to help identify a change in provenance of the mudstones of the upper

Ordovician strata.

3.2 Well Logs

Geophysical well logs record the either natural or induced response of different attributes of a formation. This study will use the data from gamma-ray, density, neutron, photoelectric, and resistivity logs. The gamma-ray log is a record of the natural radioactivity produced by the decay of potassium, uranium, and thorium near the borehole (Table 1;

Schlumberger, 2009). Spectral gamma logs can measure the contribution of each of the three elements to the total signal. Gamma-ray logs typically have higher values in shale sections compared to those composed of sandstone or carbonate (Schlumberger, 2009). The amount of natural radioactivity varies substantially between gray and black shales because the latter have a high uranium contents (Boyce, 2010). Cross-plotting the amount of thorium and potassium in shales can be used to determine clay mineral species (Adams and Weaver, 1958; Boyce, 2010;

Schlumberger, 2009).

31

Figure 3.1: Locations of wells that penetrate the Trenton Limestone within the study area. The black lines adjacent to some of the wells are horizontal wells in the study area.

32

The density log records the response of rocks near the borehole to gamma ray bombardment by the density tool. The amount of radiation returned is proportional to the electron density of the material (Schlumberger, 2009). The source delivers gamma rays into the formation that undergo scattering as a function of electron density. The tool responds to small variations in electron density, however the maximum amplitude of the deflections is only obtained from beds more than 60 cm thick (Quirein et al., 1982; Schlumberger, 2009). The density log is calibrated to a standard (either limestone or sandstone) with known water content. The electron density is used to calculate a bulk grain density, which can then be used to calculate bulk porosity (Schlumberger, 2009).

The photoelectric log records low energy gamma radiation by the formulation in units of barns/electron, it is a supplementary measurement by the density logging tools (Schlumberger,

2009). The logged value is a direct function of the aggregate atomic number of the elements in the formation, making it the best indicator of lithology (Schlumberger, 2009).

Neutron logs measure the hydrogen ion concentrations in a formation (Schlumberger,

2009). A lower neutron log reading indicates abundant formation hydrogen (Schlumberger,

2009). The neutron log measures energy loss when neutrons emitted from the tool collide with other particles in the formation. The maximum energy loss during a neutron collision occurs when a neutron collides with a particle of equal mass, that is, a hydrogen atom (Schlumberger,

2009).

The resistivity of a unit is a function of the rock type, the pore geometry, and the fluids within the pores (Schlumberger, 2009). Hydrocarbons, rock, and fresh water are all insulators that are nonconductive, yet salt water is a conductor and has a low resistivity. The 33 measurement of resistivity is a measurement of the amount (and salinity) of the formation connate water (Schlumberger, 2009).

The data were being assembled in a GeoGraphix database by Eastern Mountain Fuel and include the responses of gamma ray, density, neutron, photoelectric, and resistivity logs as well as the formation tops of each stratigraphic unit from approximately 300 oil and gas wells in

Guernsey, Muskingum, Morgan, and Noble counties. The spectral gamma ray from the Wells-

Crum in Guernsey County was used for the estimate of clay mineralogy throughout the section because it was the most complete record. XRD data from cores spanning the interval from the

Trenton to the Utica Shale (LePage and Wells-Crum) are used in the analysis of the vertical distribution of clay minerals. The LePage Unit #1 well was selected as the type well for the study, because it is located at the center of the study area. After correlation of the formation tops, east-west and north-south cross sections were constructed. Isopach and structure maps were created using the assigned formation tops within the GeoGraphix database.

3.3 XRD Data

Mineralogy data collected by X-ray Diffraction (XRD) analysis done at Weatherford labs and provided courtesy of Artex Oil Company and Eastern Mountain Fuel are available from core in two wells within the study area. One well within the study area has both XRD and spectral gamma ray data (Trenton to Queenston). The data were collected at five-foot intervals in all the wells and provide percentages of clays, carbonates, and other minerals.

The bulk mineralogy of the core samples was determined using the XRD and SEM technique by Weatherford Labs in Houston, Texas. The Rietveld XRD method was used for the

XRD data collection, which is more accurate when dealing with high amounts of clay minerals 34 and silicates. The data were reported in three main categories clays, carbonates, and other minerals. The clay data includes abundances of chlorite, kaolinite, and illite. The carbonate minerals are reported as a percentage of calcite, dolomite, and siderite. The carbonate content was further evaluated by using insoluble-residue analysis to determine how relative amounts of carbonate minerals (calcite and dolomite) vary throughout the cored interval. The other minerals present include the percent quartz, potassium feldspar, plagioclase, pyrite, apatite, marcasite, anhydrite, and total organic carbon (TOC) (Figure 3.2).

3.4 Well-log Evaluation of Mineralogy

The well log responses from the cored intervals in each well were compared to the XRD mineralogy data. First, the gamma ray log values and spectral gamma ray values (K, Th, and U) were averaged. The gamma log measures the natural radioactivity of a formation and is useful when determining clay mineral species (Quirein et al., 1982). The gamma tool receives 75% of the signal within 30cm of the borehole and has an approximate vertical resolution of 60 cm (2 feet) (Schlumberger, 2009). The core was sampled every 5 feet with gamma readings taken every 0.25 feet on the LePage #1 (Figure 3.4). The gamma responses between one foot above and one foot below core point were calculated using a weighted average in which a total of

53.5% of the gamma value occurs within 0.5 feet core point depth (Figure 3.4). Secondly, cross- plots of thorium and potassium data from the spectral gamma-ray logs were constructed and compared to the mineralogy fields proposed by Quirein et al. (1982) (Figure 3.3). Cross- plots were created in order to determine the origin of the clays (detrital vs. authigenic). Plots were made comparing the percentages the bulk weight of the XRD data and spectral gamma-ray data.

35

Figure 3.2: A sample of the XRD data of XRD data in percent of total mineral abundance from a portion of the cored interval within the Kope and Utica formations. The depths intervals are in feet.

36

Figure 3.3: A cross-plot of Thorium and Potassium values from a spectral gamma-ray tool used to interpret the type of clay minerals present (after Quirein et al., 1982).

37

Figure 3.4: An example of the methodology used to correct the total gamma-ray value and spectral gamma (K, Th, and U) value for the effects of adjacent rock. 75% of the gamma ray signal comes from 30 cm. The sum of the values on either side of the depth in question (a total of 1 foot) = 53.5% of the signal. The values for the gamma ray response at 1.25 feet on either side of the center are multiplied by 13.375%, etc. The final value for the total gamma-ray is the sum of the corrected values.

38

Once the clay plots were created, sequence stratigraphic concepts were applied to the

3rd-order sequence in order to evaluate the data from the Trenton to the top of the Queenston.

An in-depth understanding of the sequence stratigraphy will allow for a better interpretation of the provenance change. The flow chart seen in Figure 3.5 outlines the steps taken to reach the conclusions and Table 3.1 summarizes the different logs that were run on all the wells.

39

Figure 3.5: A flow chart to outlining the steps in the research.

40

Table 3.1: Summary of the different well logs and what each log measures, also the vertical resolution and depth of investigation of each log (Schlumberger, 2009).

41

CHAPTER 4: RESULTS

4.1 Introduction

The interpretation of lithofacies from the core and log data, the correlation of the lithostratigraphic tops, the construction of isopach maps and cross-plots of the core XRD data, and the spectral gamma-ray data inferred mineralogy of the remaining mudstone section all produced data sets with varying spatial and temporal resolutions.

4.2 Correlation of Lithofacies

Lithofacies are bodies of rocks that are characterized by a combination of lithology, physical properties, and biological structures that can be used to differentiate them from the surrounding and juxtaposing rock (Reading, 1996). In the subsurface core provides the most accurate data to describe and interpret the vertical succession of lithofacies and facies associations. Where core are not avail be cuttings can be used to determine general changes in lithology and formation tops. In the absence of either data set, well logs are used to interpret lithology and grain size. Because there was no direct access to core for this study the subdivision of the section is based on interpretations of well logs.

The transition from the low gamma ray values characteristic of the Trenton Limestone to the higher gamma values of the Point Pleasant Formation is identified as the Logana Member of the Trenton (Patchen et al., 2015). The Logana Member is a mixture of carbonate and argillaceous sediment that has a higher porosity than the underlying limestone. The top of the

Logana is placed at the top of the highest carbonate below the predominantly shale interval of the Point Pleasant Formation. The Point Pleasant Formation is subdivided into three informal 42 zones on gamma ray logs through variations in the amount of carbonate and siliciclastic mudstone. The top of zone 1 in the LePage well (Figure 4.1), consists of 5.8 m (19 feet) of siliciclastic mudstone capped by a thin carbonate bed. Zone 1 is the most organic-rich interval with the highest porosity. The second zone is composed of 14.3 m (47 feet) of siliciclastic mudstone with a lower organic content and higher gamma-ray values then zone 1 (Figure 4.1).

The clays within the second zone include Illite, chlorite, mixed-layered clays, and kaolinite. The amount of chlorite significantly increases from the top of zone 1 into zone 2 and the Utica

Formations. Zone 3 is composed of a limestone bed that is 2.4 m (8 feet) thick that marks the top of the Point Pleasant Formation. Figure 4.2 gives the averages of the mineralogy from XRD analyses within the cored interval of the LePage well in Muskingum County.

The Utica Formation lies directly upon the Point Pleasant Formation and contains the highest gamma-ray values in the study interval. XRD data from core show that the main clay minerals are Illite, chlorite, mixed-layered clays, and kaolinite (Figure 4.2). The top of the Utica was placed at a significant decrease in gamma-ray values that could be identified in all the wells across the study area approximately 34 m (110 feet) above the top of the Point Pleasant.

The interval between the top of the Utica and the base of the Queenston shale was subdivided into four informal units termed Mud A – D. The units vary in average thickness from

43- 82 m (140-270 feet; Table 4.1) and each is characterized by an overall decrease in gamma- ray response up section. The top of each unit is characterized by an overall decrease in gamma- ray response up section. The top of each unit is placed were the gamma-ray value rapidly increases. All the tops could be traced across the study interval. The decrease in gamma-ray up-section in each unit is a result of an increase in the amount of carbonate present also noted on the PE (photoelectric) log (Table 3.1). The interval termed Mud A is likely equivalent to the 43

Figure 4.1: Point Pleasant interval from the LePage Well (Muskingum Co. Zone 1= base of the Point Pleasant with the highest organic content. Zone 2= Siliciclastic interval of the upper Point Pleasant. Zone 3= carbonate bed used to pick the top of the Point Pleasant throughout the study area.

44

Figure 4.2: The average of the most common minerals in the XRD analyses of samples from the LePage Well. There is a general increase in the clay minerals (chlorite, Illite, mixed-layered clays, and kaolinite) and quartz of the cored interval up section. Calcite, on the other hand, decreases in the cored interval up section.

45

Table 4.1: The table displays the average thickness of the lithostratigraphic data from the study area. Each figure corresponds to its equivalent isopach map.

.

46

Kope Formation based on the minute changes in the gamma-ray values directly above the top of the Utica Shale (Patchen et al., 2015).

The Queenston Shale marks the top of the study interval. In cuttings, the top of this unit is easily recognizable by the appearance of red mudstones, however recognizing the base of the unit is more equivocal. In the study, the top of Mud D (base of Queenston) was picked where the largest decrease in gamma was observed. The base of the Queenston has the thickest zone of carbonate beds that could be correlated throughout all four counties within the study area.

4.3 Isopach Maps

The correlation of the study interval between 300 wells allowed the construction of isopach maps of each interval from the Point Pleasant to the Queenston (Figures 4.3-4.9). A total isopach interval (Figure 4.10) was also constructed. The maps of the Point Pleasant to Mud

B and the map of Mud D show variations in thickness but no overall direction in trend. The

Queenston isopach, and to a lesser extent the Mud C map, show an eastward thickening. This overall lack of trend in thickness variations is marked in contrast to the total isopach map, which shows a clear increase in thickness to the east (Figure 4.10). The orientations of the trends present in five of the maps are shown in figure 4.11 a and b. The trends vary slightly but average 94°. A compilation of the orientations of structural and depositional features that might impact the thickness of the units within the study area is illustrated in (Figure 4.12).

47

Figure 4.3: The isopach map (in feet) of the total Point Pleasant thickness shows no consistent trends. The contour interval is 10 feet. 48

Figure 4.4: The isopach map (in feet) of the Utica Shale within the study area shows no consistent thickening trend. The contour interval is 10 feet.

49

Figure 4.5: The isopach map (in feet) of the Mud A interval shows no consistent trends in thickness. The contour interval is 10 feet.

50

Figure 4.6: The isopach map (in feet) of the Mud B interval shows no consistent thickening trends. The contour interval is 10 feet. 51

Figure 4.7: The isopach map (in feet) of the Mud C interval. This map shows a general thickening to the northeast. The contour interval is 10 feet. 52

Figure 4.8: The Isopach map (in feet) of the Mud D interval within the study area. This map shows no consistent thickening trend. The contour interval is 10 feet.

53

Figure 4.9: The isopach map (in feet) of the Queenston Shale within the study area. This map shows a consistent thickening trend to the east. The contour interval is 10 feet.

54

Figure 4.10: Total isopach map (feet) of top Queenston to top Point Pleasant showing a consistent increase in thickness to the east. The contour interval is 25 feet.

55

Figure 4.11: A) Trends of isopach high values (thicker than average) within the study area of the Utica (U), Mud A (MA), Mud B (Mud B), Mud C (Mud C), and Mud D (Mud D). B) Trends of isopach low values (thinner than average) using the same nomenclature. The trends are nearly parallel and orthogonal for each interval in both cases.

56

Figure 4.12: Appalachian structural trends that affect isopach thickness trends within the study area.

57

4.4 Sequence Stratigraphy

The vertical changes in grain size interpreted from well logs can be used to make sequence stratigraphic interpretations of a section. A sequence is defined as a succession of conformable and genetically related beds bounded by unconformities (Posamentier and Vail,

1988; Van Wagoner et al., 1990, Posamentier and Allen, 1999; Catuneanu, 2002). Sequences are composed of systems tracts that represent deposition under certain relative sea level conditions. The spatial and vertical variation in facies that comprise the systems tracts are controlled by the amount and the rate of formation of the available space for the sediment to occupy, which is termed accommodation (Van Wagoner and Bertram, 1995). The rate of formation of accommodation depends on rates of change of eustasy, tectonic uplift or subsidence, and sediment flux into the basin.

The highest frequency change in relative sea level recognized in a section is termed a parasequence, which is defined as a progradational succession of beds bounded by marine flooding surfaces (Van Wagoner and B, 1995). Parasequences tend to occur as stacked sets from three types of patterns. A thinning and fining upward set is referred to as a retrogradational parasequence set (RPS) and forms in response to the rate of accommodation increasing faster than sediment can be supplied to a location. The sediments deposited under the transgressive regime form the Transgressive Systems Tract (TST). On well logs the TST can be recognized by increasing gamma-ray values on the left track of the log and increasing porosity recorded by sonic, neutron, density, or resistivity log on the right tack (Figure 4.13). The RPS is capped by the maximum flooding surface (MFS), which represents the maximum landward migration of the shoreline. Above the MFS is a progradational parasequence set (PPS) in which the sets coarsen and thicken upward (Van Wagoner, 1990). A PPS occurs when relative sea level is rising slowly 58

Figure 4.13: Sequence stratigraphic surface from gamma-ray and porosity logs.

59 and the influx of sediment into the system overwhelms the rate of formation of accommodation, and can form either a Lowstand or Highstand Systems Tract (Posamentier and

Allen, 1999). In either case the well log signature is similar; a decrease in gamma-ray values up section accompanied by a decrease in porosity (Figure 4.13). Within foreland basins the

Lowstand System Tract is seldom recognized (Patchen et al., 2015). The third stacking pattern is the Aggradational Parasequence Set (APS), that shows no vertical change in thickness or grain size, which is interpreted to mean an essentially steady state of relative sea level change.

Two different orders of sequences are present in the study section. The interval between the top of the Trenton and the top of the Queenston represent a 3rd-order sequence

(Figure 4.14). The basal sequence boundary corresponds to the top of the continuous limestone of the . The Transgressive Systems Tract (TST) extends from the lower sequence boundary to the highest gamma reading within the Utica Shale, which forms the

Maximum Flooding Surface (MFS). The basal unit of the TST (Logana Member) has increased gamma readings indicating the first deposits of siliciclastic clays. Above the Logana, the proportion of clays increases to the MFS within the Utica Shale (Figure 4.15). The Highstand

Systems Tract (HST) is much thicker than the TST and is characterized by an overall decrease in gamma-ray values to the top of the Queenston Shale.

At least six 4th-order sequences can be identified within the 3rd-order interval between the top of the Trenton Limestone and the top of the Queenston Shale. Sequence 4.1 consists of the Logana Member and zones 1 and 2 of the Point Pleasant Formation (Figure 4.15). The top of sequence 4.1 is placed at the base of the thin carbonate of zone 3. Sequence 4.2 extends from zone 3 of the Point Pleasant Formation to the top of Mud A. In this case, the MFS of 4.2 is the highest gamma reading within the Utica and is therefore coincident with the 3rd-order MFS. 60

The PPS of sequence 4.2 encompasses the Kope Formation with a sequence boundary placed at the base of a thin, regionally correlative carbonate. Sequences 4.3, 4.4, 4.5, and 4.6 correspond to the informal lithostratigraphic units of Mud B, Mud C, Mud D and the Queenston

Shale and have similar characteristics on the gamma log. The basal sequence boundaries are overlain by carbonates, which transition into mudstones in the TST. The MFS within each of the last four 4th-order sequences is thicker than the TST, however, the asymmetry of thickness between the TST and HST found in the 3rd- order sequence is only apparent in sequence 4.2.

There is also an overall trend of increasing thickness of 4th-order sequences up-section, again with the exception of sequence 4.2.

4.5 XRD Mineral Abundance Data

The XRD data obtained for this study include values for four types of clay, three types of carbonate, and seven other minerals, of which the most abundant is quartz. The most common clay mineral is Illite (1%-34%) with chlorite the next most common in most samples (0%-13%).

Kaolinite and mixed layered clays do not make up more than 8% and 9% of the samples, respectively, and averaged less than 3%. Illite occurs throughout the cored interval, however, chlorite increases significantly in abundance above 6086 feet in the LePage well. Calcite is the dominant carbonate mineral comprising up to 50% of the sample in the lower Point Pleasant

Formation. Calcite abundance drops steadily up-section in the core. The most common additional mineral is quartz, which varies from 1% - 26% with an average of 20% in the study area (Table 4.2).

61

Figure 4.14: A generalized sequence stratigraphic interpretation of the Upper Ordovician study interval. The top of the Trenton and Queenston are 3rd order sequence boundaries (Pope and Read, 1997). The Utica Shale contains the Maximum Flooding Surface (MFS). Below the Utica is the Transgressive Systems Tract (TST), while the Highstand System Tract extends from the top of the Utica to the top of the Queenston Shale. The smaller cycles within the 3rd order sequence represent 4th order changes in relative sea level. The total interval is roughly 1250 feet thick. 62

Figure 4.15: Sequence Stratigraphy of the cored interval. The Maximum Flooding Surface (3rd order) shown about 10 feet above the Point Pleasant within the Utica interval. Another Maximum Flooding Surface was identified on a 4th order scale within the middle of the Point Pleasant zone 2 (figure 4.1). The arrows represent sea level fluctuations, left (regression), and right (transgression). The cored interval is roughly 300 feet thick.

63

Table 4.2: Example of the XRD data in % bulk mineralogy from the LePage well in Muskingum County, Ohio. The increase in chlorite, illite, and quartz in the Point Pleasant Zone 2 and into the Utica suggests the interval where there is a change in provenance.

64

4. 6 Well Log Analysis of Clay Mineralogy

Variations in detrital clay mineralogy within an interval can provide valuable provenance information (Weaver, 1989). Because most of the data available are from well logs, an attempt was made to use both the spectral gamma-ray data and PEF log to determine the variations in clays present. The data from the cored interval in the LePage #1 well were compared to the XRD data of the cored interval.

The PEF log records the production of low energy gamma-rays formed in response to bombarding a formation with high-energy gamma radiation (Schlumberger, 2009). The tool is sensitive to the aggregate atomic number of the material irradiated and therefore is a very accurate indicator of quartz and calcite within a formation. The problem with using this tool to determine clay mineralogy is that many clays form in solution series with numerous elements that can be exchanged in the crystal lattice. If the exact composition of the clays in the core were known there would be an opportunity to calibrate the well log, at least locally.

Because those composition data were not available the use of the PEF log was not explored further and the subsequent analysis involves the spectral gamma-ray tool.

Comparisons of element abundance determined by comparing the spectral gamma-ray data to the total gamma ray values (Figure 4.16 a-c) show that the primary element affecting the gamma-ray response is potassium. The thorium graph has a weak positive correlation and the uranium plot shows no correlation to the gamma response. A comparison of the total gamma- ray to the amount of Illite (Figure 4.17), which is the most common clay mineral in the core

(Table A.1), has a positive correlation suggesting that Illite is responsible for much of the signal.

In order to determine if the clay minerals present in the core were detrital or authigenic 65

Figure 4.16: Three plots of elemental values from the spectral gamma-ray tool versus the total gamma-ray count from the LePage well of the cored interval. Potassium has the largest impact on the gamma-ray readings, which is consistent with the occurrence of Illite as the main clay mineral. Uranium and thorium can have a significant effect on the magnitude of the total gamma-ray value but there is no consistent trend with either element.

66

Figure 4.17: A plot of % illite versus gamma ray in the LePage well. The positive correlation indicates that most of the response of the gamma tool is from illite.

67 the clays were compared to quartz, which is most commonly detrital (Weaver, 1989). A cross- plot of quartz and Illite values from the XRD data (Figure 4.18a) has a positive correlation indicating that the Illite in the section is primarily detrital. A plot of quartz vs. chlorite (Figure

4.18b) shows a similar trend which indicates that the chlorite is also mainly detrital and therefore a potentially useful indicator of Appalachian provenance.

The final step was to use the spectral gamma ray to determine if a cross-plot of the abundance of potassium and thorium alone could be used to determine clay mineral composition (Quirein et al., 1982). Figure 4.19 a shows the results for the LePage well. The primary clay mineral in the section was identified as Illite, which is consistent with the XRD data

(Table 4.2). The scatter in the data appears to be partly a function of the amount of calcite present. However, the presence of chlorite, which the XRD data show to be the second most common clay, with ranges from 4 % - 18% was not recognized using this technique, suggesting that the log cannot determine the presence of a clay mineral if the concentrations are less than approximately 20% of the rock. The same methodology was applied to the entire Ordovician shale interval in the Wells-Crum #1 well (Figure 4.19b). The results indicate that the Ordovician mudstones are overwhelmingly illite, however, this may simply mean that the amount of other clays was below the level of resolution of the technique.

68

Figure 4.18: A) A plot of percent quartz vs. percent illite shows a strong positive correlation, which indicates that the latter is primarily detrital. The positive correlation also explains the unexpected relationship between percent quartz and gamma-ray count in figure 4.18a because as quartz increases so does the percent illite. B) The positive correlation between percent chlorite and percent illite suggests the former is detrital.

69

Figure 4.19: A) XRD mineral data from the LePage well plotted on the graph of Quirein et al., (1982). The primary clay mineral in the LePage core is illite with the scatter possibly a function of total carbonate content. B) The inferred mineralogy present in the Wells-Crum well from the base of the Point Pleasant to the top of the Queenston; using the spectral gamma-ray log. The main clay mineral interpreted to be present is illite.

70

CHAPTER 5: DISCUSSION

5.1 Introduction

A collision between tectonic plates results in two styles of flexural response that create accommodation. The load imposed by thrust sheets creates a foreland basin and the downward deflection of the overlying plate by mantle flow results in dynamic subsidence (DeCelles and

Giles, 1996). The geometry of the foreland basin is defined by the flexure of the lithosphere under the imposed thrusts (Figure 2.7). The fill within foreland basins, which varies from marine to terrestrial and can contain both (Ettensohn, 2004), is primarily in the foredeep adjacent to the thrust belt. Inboard of the peripheral bulge sedimentation in the backbulge region has a cratonic provenance while the foredeep is underfilled. Once the foredeep is overfilled there is a shift to an orogenic source for the sediment (DeCelles and Giles, 1996).

5.2 Tectonic and Eustatic Formation of Accommodation

The preservation of the thick, Late Ordovician mudstone section in the subsurface of east-central Ohio requires tectonic subsidence during the Tippecanoe Sequence. If the accommodation was a result of formation of the foredeep the isopach maps of the interval should show some evidence of a consistent thickening to the east. Instead the isopach data

(Figures 4.3-4.10) show little evidence of a regional thickening trend until the deposition of the

Queenston at the top of the interval. The orientations of the trends of the thick and thin zones suggest that these are controlled by a combination of the pre-existing structures within or near the study area, such as the Cambridge Arch and Rome Trough, the depositional and erosional trends of underlying carbonates, and the differential compaction of shale. 71

The eastward increase in thickness of the Queenston is also interpreted to be controlled by local structure rather than evidence of migration of the foredeep into the region. Muslim

(2014) showed that even with the rheology of the present, colder lithosphere, and the thrust loads of the larger Avalonian impact, the western limit of the Devonian foredeep was in central

West Virginia and Pennsylvania. Therefore the subsidence indicated by the Queenston isopach was likely related to movement on a more local structural feature, such as the Rome Trough, the western margin of which is approximately 60 miles to the east of the center of the study area.

Elimination of the subsidence of the foredeep as a source of accommodation in the study area leaves only local events overprinted by dynamic subsidence as the principle mechanism to provide the total accommodation that preserved the mudstone interval in the

Late Ordovician. The tilt of Laurentia in response to the collision of the island arc during the

Taconic Orogeny was recorded as far west as the Michigan basin during deposition of the Utica

Shale (Coakley et al., 1994).

The vertical facies variations were controlled by the rate of formation of accommodation (Van Wagoner et al., 1990), which could be a result of variations in the rate of tectonic subsidence, eustatic sea level changes, or both. The 3rd-order sequence indicates a period of rapid increase in relative sea level that culminated in the deposition of the Utica Shale

(Figure 4.14). The asymmetry in thickness between the TST and the HST in this larger sequence is typical of sequences in general. The rapid increase in rate of accommodation that generates in the TST is accompanied by a decrease in rate of sedimentation that leads to the formation of a condensed section that contains the MFS (Van Wagoner et al., 1990). The combination of total accommodation available, continued formation of accommodation, and increase in sedimentation rate after deposition of the MFS results in a thicker HST. If the pattern of 72 transgression and regression were eustatic, then the interval of maximum flooding interval may appear on compilations of Late Ordovician sea level curves (Figure 5.2). The long-term curve for

Laurentia published by Ross and Ross (1995) shows a slight increase in sea level that is more or less equivalent to the 3rd-order maxima in this study, however the maxima for the Late

Ordovician occurs during the deposition of the Mud D interval. The long-term global sea level curve of Haq and Shutter (2005) shows a continuous drawdown during the Late Ordovician. The difference between the published long-term sea level curves illustrates the effects of local tectonic accommodation. The variation between the 3rd-order relative sea level curve in this study and the published curves is also a result of tectonic accommodation, which must also have been due to dynamic subsidence.

If variations in the rate of dynamic subsidence were the only control on accommodation then the 4th-order sequences identified in this study would be expected to show patterns of thickness asymmetry within each sequence as well as a thinning and thickening of individual sequences up-section as the rate of formation of accommodation varied with time. However, that pattern is not present in general (Figure 5.1). Only sequence 4.2 is consistent with the asymmetry model and only the last three sequences show a consistent increase in thickness.

The departure from the expected thickness trends could be interpreted as the result of varying amounts of total accommodation formed by local events, such as compaction and subsidence associated with more local tectonic features or a result of eustatic variations, or a combination of both.

Six high frequency changes in relative sea level were identified in this study (Figure 4.14,

5.2). The global short-term sea level fluctuations for the same interval reported by Haq and

Schutter (2005) contain seven maxima, which is reasonable consistent with the present study. 73

The short-term variations illustrated by Ross and Ross (1995) are more complex than the global curve, but contain six maxima including one that is more or less coincident with the Utica Shale.

The close agreement on the number of short-term cycles between the two published curves and this study suggests that eustasy was the driving mechanism controlling the high frequency changes in the back-bulge zone of the Appalachian basin. The differences in the HST/TST thickness between the 4th-order sequences are interpreted to be a result of overprinting of the eustatic signal by local subsidence and compaction.

5.3 Provenance

The timing of the overfill of the Taconic foreland basin can be estimated by the change in detrital clay mineralogy in the backbulge zone (Velde, 1995). The combination of spectral gamma logs and XRD data showed that the bulk of both the illite and chlorite measured in the core was detrital. Chlorite is more commonly associated with weathering of orogenic belts; therefore, the increase in detrital chlorite near the top of the Point Pleasant Formation indicates a shift in provenance from the craton to the rising orogeny. Such an early indication of an

Appalachian sources shows that the sediment flux to the basin was able to overcome the rate of formation of accommodation prior to deposition of the Utica Shale. Ideally the vertical and spatial distribution of the changes in clay mineralogy in the study interval could have been mapped using well logs. However, comparisons of the spectral gamma-ray data to the XRD data show that clay minerals must be present in quantities greater than at least 20% to be detected using the methodology proposed by Quirein et al., (1982).

74

Figure 5.1: Model of the study area with a gamma ray profile indicating the timing of the foredeep into the study area, with sequence stratigraphic concepts added. Zone 2 of the Point Pleasant represents the initial overfill of the sediment onto the backbulge.

75

Figure 5.2: Relative sea level changes in the study interval compared to short- and long-term curves compiled for Laurentia (Ross and Ross, 1995) and globally (Haq and Schutter, 2005). The short-term variations are similar to the 4th-order fluctuations in the study area suggesting that the latter are a function of eustasy. The effects of dynamic are evident in the differences between the long-term trends and the 3rd-order curve.

76

CHAPTER 6: CONCLUSIONS

This study tested three hypotheses. The first hypothesis was that clay mineralogy can be inferred from well logs. The uses of the data from the spectral gamma-ray logs showed that the shales of the Cincinnati Group were dominantly illite. However, the comparisons of the spectral data to the XRD data from core samples illustrated that clay abundance must be more than at least 20% to be detected.

The second hypothesis was that the detrital clay mineralogy of the mudstones would show a change in provenance from the cratonic to Appalachian. The comparisons of the mineral abundances from the XRD analyses to the spectral gamma-ray showed that the changes in the illite and chlorite values within the Point Pleasant Formation were a result of the detrital influx of both.

The final hypothesis was that the accommodation for deposition was a result of dynamic subsidence. The sequences stratigraphic analysis of the well logs combined with isopach maps of the sequences in the Cincinnati Group revealed the presence of one 3rd-order and six 4th-order sequences. Comparisons of these data to published Late Ordovician long and short-term sea level changes show that not only the total accommodation but also the facies pattern with the

3rd-order sequence were controlled by dynamic subsidence. The number of 4th-order sequences is broadly consistent with those present in the published curves suggesting that eustasy was the principle driving mechanism in the formation of the short-term fluctuations in relative sea level.

Variations in the thickness of the TST and HST within the short-term sequences were likely a function of local subsidence and compaction.

77

REFERENCES

Adams, J.A., and Weaver, C.E., 1958, Thorium-to-uranium ratios as indicators of sedimentary processes: Example of concept of geochemical facies: America Association of Petroleum Geologist Bulletin, v. 42, p. 387-430.

Bergstrom, S.M., Mitchell, C.E., 1992, The Ordovician Utica Shale in the eastern mid- continental region: age, lithofacies, and region relationships, Special papers in Paleontology: A special tribute to Thomas W. Amsden: Oklahoma Geological survey Bulletin, v. 145, p. 67-89.

Biscaye, P.E., 1965, Mineralogy and sedimentation of recent deep-sea clay in the Atlantic Ocean and adjacent seas and oceans. Geological Society of America Bulletin, v. 76, p. 803-832.

Blatt, H., 1982, Sedimentary Petrology: San Francisco, Freeman, p. 564.

Bohacs, K.M., and Lazar, O.R., 2010, Sequence stratigraphy in fine-grained rocks at the field to flow unit scale: insights for correlation, mapping, and genetic controls: Applied Geoscience Conference of U.S Gulf Region, Mudstones as Unconventional /Oil Reservoirs: Houston Geological Society, Annual Meeting.

Boyce, M. L., 2010, Sub-surface stratigraphy and petrophysical analysis of the Middle Devonian interval of the central Appalachian Basin; West Virginia and Southwest Pennsylvania, PhD Thesis, West Virginia University, Morgantown, West Virginia, 158 p.

Bradley, D.C., 1989, Taconic plate kinematics as revealed by foredeep stratigraphy, Appalachian orogeny. Tectonics, v. 8, p. 1037-1049.

Brett, C. E., McLaughlin, P. I., Cornell, S. R., and Baird, G. C., 2004, Comparative sequence stratigraphy of two classic upper Ordovician successions, Trenton shelf (New York- ) and Lexington platform (Kentucky-Ohio): Implications for eustasy and local tectonism in eastern Laurentia: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 210, p. 295-329.

Castle, J.W., 1998, Appalachian basin stratigraphic response to convergent-margin structural evolution. Basin Research, v. 13, p. 397-418.

Catuneanu, O., 2002, Sequence stratigraphy of clastic systems: concepts, merits, and pitfalls: Journal of African Earth Sciences, v. 35, p. 1-43.

Coakley, B.J., Nadon, G, and Wang, W.F., Spatial variations in tectonics subsidence during Tippecanoe I in the Michigan basin: Basin Research, v. 6, p. 131-140.

Chamley, H., 1989. Clay Sedimentology: Springer, 623 p. 78

Cressman, E.R., 1973, Lithostratigraphy and depositional environments of the Lexington Limestone (Ordovician) of central Kentucky: U.S. Geological Survey Professional Paper 769, 61 p.

DeCelles, P.G., and Giles, K.N., 1996, Foreland basin systems. Basin Research, v. 8, p. 105- 129.

Decelles, P.G., Carrapa, B, Horton, B.K., and Gehrels, G.E., 2011, Cenozoic foreland basin systems in central Andes of northwestern Argentina: Implications for Andean geodynamics and modes of deformation. Tectonics, v. 30, 27 p.

Emmons, E, 1842, Geology of New York, part 2, Survey of the second geological district: New York, W. A., White and J. Visscher, 437 p.

Ettensohn, F.R., 2004, Modeling the nature and development of major Paleozoic clastic wedges in the Appalachian Basin, USA: Journal of Geodynamics, v.37, p. 657-681.

Ettensohn, F.R., 2010, Origin of Late Ordovician (mid-Mohawkian) temperate-water conditions on southeastern Laurentia: Glacial or tectonic: Geological Society of America Special Paper 466, p. 163-175.

Ettensohn, F.R., 2002, and Brett, C.E., 2002, stratigraphic evidence from the Appalachian Basin for continuation of the Taconic orogeny into early Silurian time: Physics and Chemistry of the Earth, v.27, p. 279-288.

Fisher, D.W., 1777, Correlation of the Hadrynian, and Ordovician Rocks in New York State. New York State museum Map and Chart Series 25, 75 p.

Gingele, F.X., De Deckker, P., Girault, A., Guichard, F., 2004, History of the South Java Current over the past 80 ka. Paleogeography, Paleoclimatology, Paleoecology, v. 183, p 247- 260.

Haq, B.U., and Schutter, S.R., 2008, A chronology of Paleozoic sea-level changes: Science, v. 319, p. 189-192

Holland, S. M., and Patzkowsky, M. E., 1996, Sequence stratigraphy and long-term paleoceanographic change in the middle and upper Ordovician of the eastern , in B. J. Witzke et al., eds., Paleozoic sequence stratigraphy: Views from the North American craton: Geological Society of America Special Paper 306, p. 117–129.

Holland, S. M., and Patzkowsky, M. E., 1997, Distal orogenic effects on peripheral bulge sedimentation: Middle and Upper Ordovician of the Nashville Dome: Journal of Sedimentary Research, v. 67, p. 250–263. 79

Holland, S. M. and Patzkowsky, M. E., 1998, Sequence Stratigraphic and Relative Sea-Level History of the Middle and Upper Ordovician of the Nashville Dome, Tennessee: Journal of Sedimentary Research, p. 684-699.

Huck, S., 2013, Controls on Natural Fractures in the Upper Lexington Limestone and Point Pleasant Formation: Central Ohio. Unpublished MS Thesis, Bowling Green University, Ohio, p. 113.

Kolata, D. R., Huff, W. D., and Bergstrom, S. M., 2001, The Ordovician Sebree trough: An oceanic passage to the Midcontinent United States: Geological Society of America Bulletin, v.113, p. 1067-1078.

McClain, T., 2013, Sequence Stratigraphy and Petrophysics of the late Ordovician Utica-Point Pleasant Interval in the Central Appalachian Basin, Eastern Ohio and Western Pennsylvania. Unpublished MS Thesis, West Virginia University, WV, p. 88.

McLaughlin, P. I., and Brett, C. E. 2007, Signatures of sea level rise on the carbonate margin of a late Ordovician foreland basin: A case study from the Cincinnati arch, USA: Palaios, v. 22, p. 245-267.

Muslim, Mohanad Z., M.S., August 2014, Geological Sciences, A Sequence Stratigraphic Analysis of the Berea Sandstone in Athens County, Ohio: Unpublished MS Thesis, Ohio University, 135 p.

Orton, E, 1890, First annual report of the Geological Survey of Ohio (Third Organization): Ohio Division of Geological Survey Annual Report for 1890, 323 p.

Patchen, D. G., Hickman, J. B., Harris, D. C., Drahovsal, J. A., Lake, P. D., Smith, L. B., Nyahay, R., Schulze, R., Riley, R. A., Baranoski, M. T., Wickstrom, L. H., Laughrey, C. D., Kostelnik, J., Harper, J. A., Avary, K. L., Bocan, J., Hohn, M. E., and McDowell, R., 2006, A geologic playbook for Trenton-Black River Appalachian basin exploration: U.S. Department of Energy, Final Report, 582 p.

Patchen, D.G., Carter, K.M., Hickman, J.B., Eble, C., Riley, R., Erenpreiss, M., Carter, K., Harper, J., Dunst, B., Smith, T., Cooney, M., Soeder, D., Metzger, G., Moore, J., Hohn, M., Pool, Saucer, J., 2015, Utica Shale Appalachian Exploration Consortium: U.S. Department of Energy, Final Report, 187 p.

Pope, M. C., and Read, J. F., 1997, High-resolution surface and subsurface sequence stratigraphy of Late Middle to Late Ordovician (Late Mohawkian-Cincinnatian) foreland basin rocks, Kentucky and Virginia: AAPG Bulletin, v. 81, p. 1866-1893.

80

Pope, M.C., and Read, J.F., 1998, Ordovician metrer-scale cycles: Implications for climate and eustatic fluctuations in the central Appalachians during a global greenhouse, non- glacial to glacial transition: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 138, p. 27-42.

Posamentier, H.W., and Allen, G.P., 1999, Siliciclastic Sequence Stratigraphy: Concepts and Applications: Society of Economic Paleontologists and Mineralogists Concepts in Sedimentology and Paleontology No. 7, p. 210.

Posamentier, H.W., Vail, P.R., 1988, Eustatic controls on clastic deposition. II. Sequence and systems tract model, 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: Society of Economic Paleontologists and Mineralogists Special Publication 42, p. 125-154.

Quirien, J.A., Gardner, J.S., and Watson, J.T., 1982, Combined Natural Gamma Ray Spectral/Litho-Density Measurements Applied to Complex Lithologies: Society of Petroleum Engineers Annual Meeting.

Reading, H.G., 1996, Sedimentary Environments: Processes, Facies, and Stratigraphy, Oxford University Press, p. 706.

Rogers, J., 1971, The Taconic Orogeny: Geological Society of America Bulletin, v. 82, p. 1142- 1178.

Ross, C. A., and Ross, J. R. P., 1995, North American Ordovician depositional sequences and correlations. Ordovician Odyssey, Pacific Section, Society of Economic Paleontologists and Mineralogists, 309-313 p.

Ryder, R.T., 2008, Assessment of Appalachian basin oil and gas resources: Utica-Lower Paleozoic Total Petroleum System: US Geological Survey Open-File Report, 29 p.

Schlumberger, 2009, Log Interpretation charts. SMP-7006: Schlumberger Wireline and Testing, 124, p.

Smosna, R., and Patchen, D.G., 1991, Ordovician Limestone and Shale in the Central Appalachian Basin: Early Sedimentary Response to Plate Collision. SEPM Core Workshop No. 15, Dallas, April 7, 1991, 541 p.

Swanson, V.E. 1961. Geology and Geochemistry of Uranium in Marine Black Shales. Geological Survey Professional Paper, p. 356.

Van Wagoner, J.C., Mitchum, Jr., R.M., Campion, K.M., and Rahmanian, V, D., 1990, Siliciclastic Sequence Stratigraphy in Well Logs, Core, and Outcrops: Concepts for High- resolution Correlation of Time and Facies: American Association of Petroleum Geologists Methods in Exploration Series No. 7, 55 p. 81

Van Wagoner, J.C., and Bertram, G.T., eds., 1995, Sequence Stratigraphy of Foreland Basin Deposits: Memoir, American Association of Petroleum Geologists, Tulsa, v. 64, p. 487.

Vanuxem, L, 1838, Second Annual Report of the Geological Survey of the Third District of the State of New York: New York Geological Survey Second Annual Report, p. 253-286.

Velde, B., 1995, Origin and mineralogy of clays. Clays and the Environment, Springer-Verlag, 356 p.

Weaver, C. E., 1989, Clays, Muds, and Shales. Developments in sedimentology, SEPM 44. Amsterdam: Elsevier, p. 819.

Wickstrom, L.H., Gray, J.D., and Stieglitz, R.D., 1992, Stratigraphy, structure, and Production history of the Trenton Limestone (Ordovician) and adjacent strata in northwestern Ohio: Ohio Divisions of Geological Survey, Report of Investigations No. 143, p. 78.

82

APPENDIX: A XRD Data

LePage CLAYS Carb. Mx Depth Form. Chlorite Kaolinite Illite Calcite Fe-Dol Siderite I/S* 5946 Mud A 11 4 26 2 15 1 0 5951 Mud A 13 0 26 2 21 2 0 5956 Mud A 12 0 33 3 14 1 0 5961 Mud A 13 0 34 3 11 0 0 5966 Mud A 12 0 32 3 16 1 0 5971 Mud A 12 0 30 3 14 4 0 5976 Mud A 11 1 28 3 18 3 0 5981 Utica 5 8 24 1 23 4 0 5986 Utica 18 0 29 4 11 5 0 5991 Utica 10 0 29 2 27 4 0 5996 Utica 12 0 28 3 18 5 0 6001 Utica 14 0 32 4 11 2 0 6006 Utica 11 2 27 3 16 5 0 6011 Utica 10 7 25 3 19 1 0 6016 Utica 11 0 31 6 13 1 0 6021 Utica 10 0 33 6 14 1 0 6026 Utica 11 0 28 3 17 1 0 6031 Utica 4 0 10 1 70 3 0 6036 Utica 13 0 25 3 21 4 0 6041 Utica 11 0 32 7 13 2 0 6046 Utica 9 0 26 5 21 4 0 6051 Utica 8 0 20 6 24 5 0 6056 Utica 9 0 33 8 11 1 0 6061 Utica 11 0 36 9 7 0 0 6066 Utica 10 0 33 8 8 2 0 6071 Utica 8 0 28 7 23 4 0 6076 Utica 10 1 27 7 18 3 0 6081 Utica 10 1 27 7 17 4 0 6086 Zone 3 10 1 22 5 24 5 0 6091 Zone 3 8 1 18 3 45 3 0 6096 Zone 2 3 3 15 4 35 4 0 6101 Zone 2 5 2 13 4 44 2 0 6106 Zone 2 5 2 10 3 49 2 0 6111 Zone 2 9 3 25 4 29 1 0 6116 Zone 2 8 2 21 3 27 2 0 83

6121 Zone 2 7 2 18 3 38 2 0 6126 Zone 2 9 0 18 3 38 2 0 6131 Zone 2 9 0 19 3 34 2 0 6136 Zone 2 9 0 20 3 32 2 0 6141 Zone 1 6 0 13 2 50 4 0 6146 Zone 1 0 0 3 1 82 3 0 6151 Zone 1 1 0 9 1 63 2 0 6156 Zone 1 1 0 9 1 63 2 0 6161 Logana 5 0 14 3 51 3 0 6166 Logana 0 0 1 0 92 2 0 6171 Logana 2 0 13 1 69 2 0 6176 Logana 0 0 3 1 86 2 0 6181 Logana 1 0 7 1 77 3 0

OTHER MINERALS Depth Form. Quartz K-spar Plag. Pyrite Apatite Marc. Anhy. TOC 5946 Mud A 26 3 7 2 1 0 0 0 5951 Mud A 25 2 6 2 1 0 0 0 5956 Mud A 24 2 6 2 2 0 0 1 5961 Mud A 25 2 7 4 0 0 0 1 5966 Mud A 24 2 6 2 1 0 0 1 5971 Mud A 23 2 6 2 1 1 0 1 5976 Mud A 24 1 5 3 1 0 0 1 5981 Utica 22 2 6 2 1 0 0 1 5986 Utica 22 2 5 2 1 0 0 1 5991 Utica 19 1 5 1 1 0 0 1 5996 Utica 23 1 5 2 1 0 0 1 6001 Utica 24 2 6 2 1 0 0 1 6006 Utica 23 2 5 2 2 0 0 1 6011 Utica 23 2 5 2 1 0 0 1 6016 Utica 26 2 5 2 1 0 0 1 6021 Utica 26 2 6 2 0 0 0 0 6026 Utica 26 2 7 2 1 0 0 2 6031 Utica 8 1 2 1 0 0 0 0 6036 Utica 23 2 6 2 1 0 0 0 6041 Utica 24 2 6 2 1 0 0 0 6046 Utica 25 1 5 2 1 0 0 1 6051 Utica 24 2 6 2 1 0 0 2 84

6056 Utica 27 2 6 2 1 0 0 0 6061 Utica 25 2 6 1 1 0 0 2 6066 Utica 26 3 6 1 1 0 0 2 6071 Utica 22 2 4 1 1 0 0 0 6076 Utica 22 3 6 0 1 0 0 2 6081 Utica 23 1 5 2 1 0 0 2 6086 Zone 3 22 2 5 1 1 0 0 2 6091 Zone 3 12 1 4 1 1 1 0 2 6096 Zone 2 24 2 5 1 1 1 0 2 6101 Zone 2 22 1 3 1 1 0 0 2 6106 Zone 2 19 1 3 1 1 0 0 2 6111 Zone 2 18 1 3 1 1 1 0 2 6116 Zone 2 22 2 4 1 2 0 0 2 6121 Zone 2 21 1 3 0 1 0 0 2 6126 Zone 2 19 1 3 1 0 0 0 2 6131 Zone 2 18 1 4 1 1 0 0 3 6136 Zone 2 19 1 6 1 1 0 0 4 6141 Zone 1 13 1 4 1 1 1 0 4 6146 Zone 1 4 0 2 3 1 0 0 3 6151 Zone 1 12 0 3 2 1 0 0 5 6156 Zone 1 12 0 3 2 1 0 0 5 6161 Logana 13 1 3 1 1 1 0 4 6166 Logana 1 0 1 0 0 0 3 0 6171 Logana 8 1 2 1 1 0 0 0 6176 Logana 3 0 1 0 2 0 0 0 6181 Logana 5 0 2 1 3 0 0 0

Wells- Crum CLAYS CARBONATES Depth (ft) Form. Chlorite Kaolinite Illite M/s Calcite Dolo Sid 6165.00 Mud A 10 0 32 3 12 1 0 6175.00 Mud A 12 0 32 3 14 1 0 6185.00 Mud A 10 0 20 2 35 1 0 6195.00 Mud A 10 0 35 3 14 2 0 6205.00 Mud A 11 2 19 5 23 5 0 6215.00 Mud A 11 0 32 3 17 4 0 6225.00 Utica 10 0 35 3 14 1 0 6235.00 Utica 12 0 29 3 15 6 0 6245.00 Utica 16 0 27 3 18 2 0 85

6255.00 Utica 11 0 36 3 12 1 0 6265.00 Utica 11 0 32 3 18 5 0 6275.00 Utica 10 0 34 3 16 4 0 6285.00 Utica 10 0 32 3 21 2 0 6295.00 Utica 10 0 32 5 12 5 0 6305.00 Zone 3 13 2 21 4 22 5 0 6310.00 Zone 3 1 0 4 0 86 2 0 6315.00 Zone 2 7 4 22 4 23 8 0 6320.00 Zone 2 6 3 18 2 41 1 0 6325.00 Zone 2 6 3 19 2 35 2 0 6330.00 Zone 2 5 1 33 2 22 2 0 6335.00 Zone 2 7 1 24 2 31 2 0 6340.00 Zone 2 1 0 29 2 35 2 0 6345.00 Zone 2 1 0 24 2 41 2 0 6350.00 Zone 1 1 0 19 1 49 3 0 6355.00 Zone 1 1 0 23 2 44 3 0 6360.00 Zone 1 1 0 24 2 41 4 0 6365.00 Zone 1 1 0 23 2 39 7 0 6370.00 Logana 1 0 20 1 46 3 0 6375.00 Logana 1 0 14 1 52 3 0 6380.00 Logana 1 0 23 1 40 8 0 6385.00 Logana 1 0 11 1 67 4 0 6390.00 Logana 0 0 11 1 70 3 0 6395.00 Logana 0 0 12 1 69 3 0 6400.00 Logana 0 0 1 0 93 0 0

Depth Form. Quartz K-spar Plag. Pyrite Apatite Marcasite TOC 6165.00 Mud A 25 3 8 4 1 0 0 6175.00 Mud A 27 3 7 3 1 0 0 6185.00 Mud A 23 2 8 2 1 0 0 6195.00 Mud A 25 2 7 2 0 0 0 6205.00 Mud A 20 2 6 4 2 0 1 6215.00 Mud A 22 2 5 4 1 0 1 6225.00 Utica 23 2 6 3 2 0 1 6235.00 Utica 23 2 6 2 0 0 1 6245.00 Utica 21 2 6 2 1 0 1 6255.00 Utica 24 2 6 2 2 0 1 6265.00 Utica 21 2 5 2 2 0 1 86

6275.00 Utica 23 2 6 2 0 0 0 6285.00 Utica 21 2 5 2 1 0 1 6295.00 Utica 24 2 6 2 1 0 0 6305.00 Zone 3 21 2 6 2 0 0 0 6310.00 Zone 3 4 0 2 1 0 0 0 6315.00 Zone 2 20 1 3 3 2 1 2 6320.00 Zone 2 21 1 3 2 0 0 2 6325.00 Zone 2 22 1 5 2 0 0 3 6330.00 Zone 2 24 1 4 2 1 0 2 6335.00 Zone 2 21 1 4 3 1 0 2 6340.00 Zone 2 22 1 4 1 1 0 2 6345.00 Zone 2 17 1 5 3 1 0 3 6350.00 Zone 1 14 1 5 3 1 0 3 6355.00 Zone 1 18 1 5 1 0 0 2 6360.00 Zone 1 15 1 5 2 1 1 3 6365.00 Zone 1 16 1 4 3 1 0 3 6370.00 Logana 22 0 2 2 0 0 3 6375.00 Logana 16 1 3 3 2 0 4 6380.00 Logana 16 1 3 2 2 1 2 6385.00 Logana 8 0 2 2 2 0 2 6390.00 Logana 8 1 2 1 1 0 2 6395.00 Logana 7 1 2 2 1 0 2 6400.00 Logana 1 0 1 1 0 0 3

87

APPENDIX: B WELLS

Well Name County 1 MARSHALL W R Guernsey 2 VESSELS Guernsey 3 VALENTINE Guernsey 4 SCHLABACH Guernsey 5 HOLMES LS Guernsey 6 EIKENBERRY Guernsey 7 BADERTSCHER Guernsey 8 HOLMES Guernsey 9 SCION-BARSTOW Guernsey 10 MORRIS Guernsey 11 SEI-GROVE Guernsey 12 MATHERS Guernsey 13 PERRY Guernsey 14 EIKENBERRY Guernsey 15 JACOBS Guernsey 16 JACOBS UNIT Guernsey 17 BELL UNIT Guernsey 18 E MORRIS UNIT Guernsey 19 WOLTZ Guernsey 20 SHAFFER Guernsey 21 STOVER S Guernsey 22 RINGER Guernsey 23 MORRIS-MISOCK Guernsey 24 LUBURG Guernsey 25 AULT Guernsey 26 PENNINGTON Guernsey 27 MATHERS/MACFARLAND Guernsey 28 BARBARA Guernsey 29 CARMEN Guernsey 30 MAYA Guernsey 31 WELLS-WILSON Guernsey 32 HOLMES LIMESTONE Guernsey 33 MAGIS UNIT Guernsey 34 HAUN-BRITNELL Guernsey 35 BISSONETTE UNIT Guernsey 36 RINGER-MATHERS Guernsey 37 BARTHOLOW AUTO Guernsey 88

38 ENOS UNIT Guernsey 39 MARY ANN Guernsey 40 DUNLEAVY Guernsey 41 MORGAN Guernsey 42 GRIFFIN UNIT Guernsey 43 PRM Guernsey 44 W NEILLY Guernsey 45 MAGIS UNIT Guernsey 46 GROVE UNIT Guernsey 47 CRUM-BURKE Guernsey 48 CARL MATHERS Guernsey 49 BARNES UNIT Guernsey 50 MISOCK Guernsey 51 HIGGINS Guernsey 52 DEVCO UNIT Guernsey 53 HUDSON UNIT Guernsey 54 MISOCK Guernsey 55 RINGER-NEILLEY Guernsey 56 ARCHER-SHEGOG Guernsey 57 MCFARLAND-NEILLEY Guernsey 58 NEILLEY B & M Guernsey 59 WELLS-CRUM Guernsey 60 ELLIOT Guernsey 61 MCCANCE-KNOWLTON Guernsey 62 ARCHER-SHEGOG Guernsey 63 MITCHELL Guernsey 64 MITCHELL-RYAN Guernsey 65 FREC GUER-SPENCER Guernsey 66 HANN Morgan 67 MAUTZ-CATON UNIT Morgan 68 DETWEILER UNIT Morgan 69 SMITH FAMILY FARM UNIT Morgan 70 CAMPBELL Morgan 71 D K HOOK Morgan 72 HANCHER Morgan 73 J PALMER Morgan 74 ST HARPER Morgan 75 WILLEY-BINION-HOOK Morgan 76 HESS Morgan 77 C B WILLEY Morgan 89

78 OWEN REED Morgan 79 WORK Morgan 80 VAN HORN Morgan 81 HALL Morgan 82 BARKMAN Morgan 83 WEAVER UNIT Morgan 84 BANKES Morgan 85 WEAVER Morgan 86 LAPP-LEMON Morgan 87 PATTERSON UNIT Morgan 88 DAYSPRING UNIT Morgan 89 SIGMAN VINCENT Morgan 90 PATTERSON-REED UNIT Morgan 91 BARKMAN UNIT Morgan WOODWARD-CAMPBELL 92 UNIT Morgan 93 SMITH-FRALEY UNIT Morgan 94 STOVER UNIT Morgan 95 M & E MILLER Morgan 96 LAPP UNIT Morgan 97 L K OSBORN Morgan 98 LEEDOM Muskingum 99 CON COAL Muskingum 100 DEARTH Muskingum 101 KENNEDY Muskingum 102 BLANCETT Muskingum 103 MAUTZ Muskingum 104 BARNETT Muskingum 105 JOHNSTON Muskingum 106 WELCH/KING Muskingum 107 JOHNSTON/FUSNER Muskingum 108 ANDERSON Muskingum 109 MELL Muskingum 110 HOBBS Muskingum 111 GORMLEY Muskingum 112 MILLRUN Muskingum 113 PARR Muskingum 114 STRATTON Muskingum 115 BOOTH Muskingum 116 CHARLES Muskingum 90

117 DEW-BICKLEY Muskingum 118 VANDENBARK Muskingum 119 HARDING Muskingum 120 DEITRICK Muskingum 121 FARMS Muskingum 122 KELLY Muskingum 123 C RINK Muskingum 124 OHIO POWER Muskingum 125 OHIO POWER Muskingum 126 MYERS CHARLES J & JOAN Muskingum 127 WANDA Muskingum 128 WANDA Muskingum 129 WANDA Muskingum 130 CASEY Muskingum 131 ANDERSON Muskingum 132 C GOSS Muskingum 133 MILLER Muskingum 134 WILSON Muskingum 135 WILSON Muskingum 136 VICKI Muskingum 137 DANIELLE Muskingum 138 KNICELY Muskingum 139 WATSON Muskingum 140 COSSEL Muskingum 141 VICKERS Muskingum 142 BRANDY Muskingum 143 KATHY Muskingum 144 CARRIE Muskingum 145 SALLY Muskingum 146 KAYLA Muskingum 147 CINDY Muskingum 148 BELL Muskingum 149 MARLING Muskingum 150 HUGGINS R Muskingum 151 ROSE Muskingum 152 BETTINA Muskingum 153 WENDY Muskingum 154 EOC-HUFFMAN Muskingum 155 DEARINGER Muskingum 156 HEATHER Muskingum 91

157 TONYA Muskingum 158 THOMAS Muskingum 159 WANDA Muskingum 160 TINA Muskingum 161 KING Muskingum 162 CHARLOTTE Muskingum 163 LEAH Muskingum 164 MARSHALL Muskingum 165 TROY TOM Muskingum 166 EOC-DANIEL Muskingum 167 EOC-HOUK Muskingum 168 TINA Muskingum 169 COPA Muskingum 170 BOWDEN Muskingum 171 FRANCE Muskingum 172 LYNN Muskingum 173 BIEDENBACH Muskingum 174 APRIL Muskingum 175 WORSTALL Muskingum 176 JUNE Muskingum 177 RASOR Muskingum 178 COLLOPY Muskingum 179 MONICA Muskingum 180 SUGGETT Muskingum 181 ECK Muskingum 182 VINSEL Muskingum 183 RHONDA Muskingum 184 GIERKE Muskingum 185 GIERKE Muskingum 186 CALDWELL-FRENCH Muskingum 187 LUBURGH UNIT Muskingum 188 MUSK RIVER Muskingum 189 WOERNER-HITTLE Muskingum 190 ERIN Muskingum 191 RANDI Muskingum 192 DENISE Muskingum 193 HUHN Muskingum 194 GIBSON TRUST Muskingum 195 CAMERON Muskingum 196 SIERRA Muskingum 92

197 VANDENBARK Muskingum 198 EVE Muskingum 199 RANDI Muskingum 200 LAPP/RADCLIFFE Muskingum 201 PRAIRIE FORK Muskingum 202 ERMA Muskingum 203 CHARLENE Muskingum 204 SEI-MYERS Muskingum 205 SEI-MCCUTCHEON Muskingum 206 SEI-HOUK Muskingum 207 ECK Muskingum 208 HEIDI Muskingum 209 ERICA Muskingum 210 GAY Muskingum 211 KIRSCH UNIT Muskingum 212 CONSOL COAL CR Muskingum 213 ELLEN Muskingum 214 ELLEN Muskingum 215 JERI Muskingum 216 KREPS UNIT Muskingum 217 DEBBIE Muskingum 218 MARY JANE Muskingum 219 HOCK Muskingum 220 MITCHELL Muskingum 221 MCCORMICK-WATSON Muskingum 222 NETTIE Muskingum 223 NETTIE Muskingum 224 JOY Muskingum 225 JOY Muskingum 226 TRACIE Muskingum 227 JOAN Muskingum 228 MYERS UNIT Muskingum 229 PAUL UNIT Muskingum 230 DEAL Muskingum 231 RUBY UNIT Muskingum 232 HERNDON Muskingum 233 DEAL Muskingum 234 ELLIOT Muskingum 235 BOOTH-TOWNING Muskingum 236 KLIES C Muskingum 93

237 CONSOL COAL Muskingum 238 T MUCUTCHEON Muskingum 239 WATTS Muskingum 240 SEI-WISE UNIT Muskingum 241 SHANNON Muskingum 242 FRAZIER UNIT Muskingum 243 HATFIELD UNIT Muskingum 244 KATHLEEN Muskingum 245 KAY Muskingum 246 KELLY Muskingum 247 LOUISE Muskingum 248 SHANNON Muskingum 249 TRACI Muskingum 250 LOUISE Muskingum 251 PATTY Muskingum 252 EMERSON Muskingum 253 MISSY Muskingum 254 TRACI Muskingum 255 MEYER-REED UNIT Muskingum 256 KENNISON-FENTON Muskingum 257 SHARON Muskingum 258 MENZIE-FRICK Muskingum 259 TOM Muskingum 260 JOHNSON UNIT Muskingum 261 APPERSON UNIT Muskingum 262 MENZIE-FRICK UNIT Muskingum 263 DAVY-DOWNING UNIT Muskingum 264 BEATTY-DAVY UNIT Muskingum 265 WILKINS-DOWNING UNIT Muskingum 266 KENNEDY-FRECO UNIT Muskingum 267 WATTS-WILKINS Muskingum 268 SHURTER-PARMER Muskingum 269 MITCHELL RALPH Muskingum 270 BEST-WILLIS UNIT Muskingum 271 ALEXANDER-HARRIS Muskingum 272 REED-ADAMS Muskingum 273 BAILEY Muskingum 274 HOLSKEY Muskingum 275 OKEY-ECK Muskingum 276 GALLOGLY-PERKINS Muskingum 94

277 RITTENHOUSE-PERKINS Muskingum 278 FRECO-LEPAGE Muskingum 279 DAVY-DOWNING Muskingum 280 HARMON-FENTON Muskingum 281 MEYER KIMBERLY Muskingum 282 BYRNS JAMES Muskingum 283 WELCH-SEARS Muskingum 284 DAVY-VINEYARD Muskingum 285 LEPAGE K Muskingum 286 EUCKER D & N Muskingum 287 WONDERLAND-KELSO Muskingum 288 BYRNS-KELSO Muskingum 289 DICKSON-COX Muskingum 290 FRECO-FRICK UNIT Muskingum 291 SCHELL-RIGGENBACH Muskingum 292 FREC MEIGS Muskingum 293 L K OSBORN Muskingum 294 DANFORD Noble 295 BANIA Noble 296 BANIA Noble 297 ARCHIBALD Noble 298 OHIO POWER Noble 299 CENTRAL OHIO COAL Noble 300 VAN FLEET Noble 301 MURPHY Noble 302 STONEKING Noble 303 COOPER UNIT Noble 304 COOPER UNIT Noble 305 MILEY GAS CO UNIT Noble 306 MCELROY Noble 307 MILEY Noble 308 FERC NOBL OLIVE Noble 309 FREC NOBL BROOKFIELD Noble

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

Thesis and Dissertation Services ! !