CALIFORNIA STATE UNIVERSITY, NORTHRIDGE

GEOLOGIC INTERPRETATIONS OF SEASAT-A RADAR

IMAGES AND LANDSAT MSS IMAGESOF A PORTION

OF THE SOUTHERN APPALACHIAN PLATEAU:

VIRGINIA, KENTUCKY, WEST VIRGINIA

A thesis submitted in partial satisfaction of the requirements for the degree of Master of Science

in

Geology

by

Judd Muskat

MAY, 1983 The Thesis of Judd Muskat is approved:

John P. Ford (Date)

George C. ~nne (Date) v

Herbert G. Adams; Chairman (Date)

California State University, Northridge

ii DEDICATION AND ACKNOWLEDGEMENTS

This thesis is dedicated to my wife Barbara, and our two children, Joseph Jack and Jessica Ruth.

I would like to thank my thesis advisor Dr. Herbert G. Adams for providing guidance and support throughout the duration of this study. I thank the members of the Thesis Committee; Dr. George C. Dunne for constructive comments and revisions of this text; and Dr. John P. Ford for suggesting this project and providing invaluable advice and consultations. I want to thank Dr. Charles Elachi for making the facilities of the Jet Propulsion Laboratory available to me, M. L Daily, R. G. Blom, and M. L. Bryan for their advice and consultations, Ms. Annie L. Holmes for help in acquiring the imagery used in this study, and Ms. Susan Conrow who aided in preparation of__ the graphics.

The research described in this paper was performed at the Jet Propulsion Laboratory, California Institute of Technology, under contract with the National Aeronautics and Space Administration.

iii TABLE OF CONTENTS

TITLE PAGE • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • i

APPROVAL PAGE • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • ii

DEDICATION AND ACKNOWKLEDGEMENTS • • • • • • • • • • • • • • • • • • • • • iii

TABLE OF CONTENTS • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • iv

LIST OF FIGURES ...... v

ABSTRACT ••••••••••••••••••••••••••••••••••••••••••••••••••• viii

INTRODUCTION 1

Previous Works • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 8

METHODS •••••••••••••••••••••••••••••••••••••••••••••••••••• 10

Radar Imagery and Seasat-A 21

Landsat Series 26

ANALYSIS ...... 27

Drainage 27

Surface Lithology and _ Image Texture ...... 42

Structore .•••.•••.•••••••••••••••.••.••••••••••••.••••••.••• 47_

Indian Creek Fault •••.••••••.•.••••••••.••••••••••••••••••• 58

Guest Mountain Fault 60

Other Possible Faults 73

Lineaments 80

SUMMARY 94 REFERENCES CITED ...... 96 APPENDIX United States Geological Survey Maps Used for this Study ••••••••••••••••••••••••••••••••••••• 100

iv LIST OF FIGURES

1. Area Geologic Map ...... • ...... 3

2. Map of Appalachian Tectonic Divisions • • • • • • • • • • • • • • • • • • • • • 5

3. Cross Section Through Appalachian Plateau • • • • • • • • • • • • • • • • • 7

4. Seasat-A SAR Image Radar Look Direction N 67.70W ••••••••••••••••••••••••• 12

5. Seasat-A SAR Image, Radar ·Look Direction N 67.70E ••••••••••••••••••••••••••••••• 14

6. Landsat MSS Winter Subscene • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 16

7. Landsat MSS Spring Subscene ••••••••••••••••••••••••••••• 18

8A. Seasat-A SAR Image of Pine Mountain Thrust Plate, Radar Look Direction N 67 .7ow • • . . . • . . • . . . . . • ...... • . • . • . . . • • • • • • • . • . . • • • • . 20

8B. Seasat-A SAR Image of Pine Mountain Thrust ~late, Radar Look Direction N 67 .70E • . . . • • • ...... • • • • • • . . • • • . • • • • . • • • . • • . • • . • • • • . 20

8C. Landsat MSS Image of Pine Mountain Thrust Plate ...... • . . . . . • ...... • . . . . . • . . 20

8D. Geologic Structure Interpretation Map From Seasat-A SAR and Landsat MSS Images of Pine Mounain Thrust Plate • • • • • • • • • • • • • • • • • • • • • 20

9 Schematic Representation of Interaction Between Radar and Various Surfaces • • • • • • • • • • • • • • • • • • • • • 23

10. Generalized Interpretation Map of Seasat-A SAR, Radar Look N 67. ?OW • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 29

11 Drainage Map of Study Area from Blue Lines on U.S. Geological Survey 1:250,000 Series Topographic Maps • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 31

12. Drainage Interpretation Map of Seasat-A SAR, Radar Look N 6 7. ?OW • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 33

13. Drainage Interpretation Map of Seasat-A SAR, Radar Look N 67 .70E • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 35

v 14. Drainage Interpretation Map of Winter Landsat MSS Subscene • . • . . • • • . • . • . . . . . • • • . • . • ...... • . . 37

15. Drainage Interpretation Map of Spring Landsat MSS Subscene • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 39

16. Outcrop of Mississippian Limestone and Shale Exposed Along the Crest of Pine Mountain ••••••••••.•••••••••••••••••••••••••••.. 43

17. Typical Appalachian Plateau Stratigraphic Section ...... • . • ...... 45

I 18. Geologic Correlation Chart • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 46

19. Surface Texture Patterns, Seasat-A SAR Image Radar Look N 67.70W •••••••••••••••••••••••••••• 49

20. Surface Texture Patterns, Seasat-A SAR Image Radar Look N 67.70E ••••••••••••••••••••••••••••• 51

21. Surface Texture Patterns, Landsat MSS Winter Subscene • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 53

22 Schematic _Map of Geologic Structures North of Pine Mountain Thrust Plate • • • • • • • • • • • • • • • • • • • • • 55

23. Tilted Shale Beds Along the Trace of the Indian Creek Fault • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 59

24. Slickensides Exposed Along the Trace of the Indian Creek FAult • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 62 - 25. Slickensides Exposed Along the Trace of the Indian Creek Fault • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 64

26. Slickensides Exposed Along the Trace of the Indian Creek Fault •••••••••••••••••••••••••••••••••- 66

27. Guest Mountain Fault Exposed in a Roadcut • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • . • • • • • • • • • • • 68

28. Line Drawing of an Offset Coal Bed Exposed Along Guest Mountain Fa:ul t ...... 70

29. Gouge Zone of Guest Mountain Fault • • • • • • • • • • • • • • • • • • • • • • • 72

30. Structure Contours Drawn on Top of the Ohio Shale ...... 7 5

vi 31. Residual Total Intensity Aeromagnetic Map • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 77

32. Simple Bouguer Gravity Map • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 79

33 Lineament Interpretation Map, Seasat-A SAR, Radar Look N 67 .70W • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 82

34. Lineament Interpretation Map, Seasat-A SAR, Radar Look N 67 .70E • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 84

35. Lineament Interpretation Map, Landsat MSS Winter Subscene • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 86 I 36A. Rose Diagram Showing Lineament Frequency Per Azimuth Class, Seasat-A SAR, Radar Look N 67.70W .••..••..••.•.•...•.••.••••.••••.. 87

36B. Rose Diagram Showing Lineament Frequency Per Azimuth Class, Landsat MSS Winter Subscene • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • • 87

37. Lineaments from Seasat-A SAR, Radar Look N 67 .7ow Which Are Not Seen on Landsat MSS Winter Subscene • • • • • • • • • • • • • • • • • • • • • • • • 90

38. Lineaments Mapped from Landsat MSS Winter Subscene Which Are Not Seen on Seasat-A SAR, Radar Look N67.7ow ...... ••...... •...... 92

vii ABSTRACT

GEOLOGIC INTERPRETATIONS OF SEASAT-A RADAR IMAGES AND

LANDSAT MSS IMAGES OF A PORTION OF THE SOUTHERN APPALACHIAN

PLATEAU: VIRGINIA, KENTUCKY, WEST VIRGINIA

by

Judd Muskat

Master of Science in Geology

Seasat-A, a NASA oceanographic research acquired high

resolution synthetic aperture radar (SA.R) images of a portion of the

Appalachian Plateau during its brief mission in the summer of 1978.

Interpretation maps of surface drainage patterns, surface texture patterns

and surface lineament patterns were made from the Seasat-A SAR images

of southwestern Virginia, eastern Kentucky, and southwestern West

Virginia. These maps are compared with similar data derived from

Landsat multispectral scanner (MSS) images of the same area. In regions ' of low topographic relief the Seasat-A SAR images are superior to the

Landsat MSS images for drainage and lineament mapping. This is due to

the higher image resolution of the Seasat radar imagery;

viii also the inherent illumination geometry of the Seasat-A SAR provides better detection of subtle topographic features. In. areas of high topographic relief, the Seasat-A SAR images are distorted because of the radar layover effect, but still the Seasat-A SAR is as good as the Landsat

MSS images for geologic interpretations. Field checking of a prominent lineament located on the Pine Mountain thrust plate of southwestern

Virginia has confirmed the existence of a previously unreported fault 16 km long, with both strike-slip and vertical displacement. A previously unmapped thrust fault was also recognized in the field. These two faults are named herein. The thrust fault shows 30m of stratigraphic throw.

ix INTRODUCTION

This thesis concerns geologic interpretations of orbital imagery acquired by the Seasat-A and Landsat satellites over portions of southeastern Kentucky, southwestern Virginia, and southwestern West

Virginia. The purpose of this study is to evaluate the use of spaceborne radar imagery as a tool for geologic reconnaissance in areas of limited or difficult access due to rugged terrain and dense vegetation cover, or which may be hidden from other systems by perennial cover. The southern Appalachian Plateau was chosen as a study area because of the rugged and steep terrain, the near total vegetation cover, the large volume of background geologic literature available, the interesting and complex nature of the local geologic structure, and the availability of good quality Seasat-A radar images.

The study area covers approximately 11,500 km2 and encompasses the transition zone of the southern Appalachian Plateau with the Valley and Ridge geomorphic province to the east (Figure 1). The local terrain is

maturely_ dissected with slopes that range mostly from about 160 to 260.

The relatively small areas of level surfaces occur in the flood plains of the major streams. Irregular winding ridges and narrow V-shaped valleys occupy approximately equal portions of the landscape. Maximum relief is

600 m between the highest peak (Pine Mountain) and the valley floors.

The Appalachian Plateau and Valley and Ridge provinces collectively are incorporated in the Paleozoic Appalachian Basin. The

Appalachian Basin comprises a small part of the foreland trough which extends the full length of the Appalachian orogenic (mobile) belt (Figure

2).

1 2

Figure 1. Area geologic map (from Harris and Milici, 1977). On this figure the study area is outlined by hachures (central portion of figure). Also highlighted are the paths of the two orbits (revolutions 694 and 874) from which the Seasat-A SAR images used in this study were obtained (Figures 5, 6, 8A, and 8B). _,

REV ( ) 6 9 4 I I j INDIANA

I I I KENTUCKYj I I

3 a•

PALEOZOIC METAMORPHIC and IGNEOUS ROCKS

IDSOCI DEVONIAN lhmugh ORDOVICIAN ROCKS IDSO)

SIT.URIAN and ORDOVTOAN ROCI(S ISO)

SILURIAN lhrough CAMBRIAN ROCKS tSOCI

"""""""'QUAimmS [c:-"kl PRECAMBRIAN METAMORPHIC IGNEOUSall\ ROCKS 82° ""'"" Eoult ---~------~---

4

I

Figure 2. Regional map showing the several Appalachian tectonic divisions (from Dott and Batten, 1971). Cross section A-A' passes through the study area and is presented in Figure 3. 5

APPALACHIAN TECTONIC DIVISIONS

Large folds

Steep faults

low angle thrust faults

D Late Triassic rocks in fault basins

- Low grade metamorphic rocks 0

Ld High grade metamorphic rocks

I ~ Paleozoic granitic plutons ~

\

'. .\ '-·-., // ·...... -·"...... / \ ...... -. \ / \ ...... \ \ ' ...... ~.,_ "'< ....- '

\ ,. \ \ \

\ "\ \ \ \ ' ) 01------4~0TOI '• ______B,QOkm. , 0 250 500 mi. 6

The southern Appalachian Plateau is underlain by a sequence of unmetamorphosed Paleozoic miogeosynclinal carbonates and clastics which lie upon Precambrian crystalline basement. The dominant structures of the southern Appalachian Plateau and Valley and Ridge provinces were formed during the late Paleozoic Allegehenian orogeny.

Several workers have concluded that gravity was the driving mechanism of that deformation (Gwinn, 196/i.; Miller, 1973; Dennison, 1974; Milici,

1975; and Chappel, 1978). However, most workers conclude that deformation was of a "thin-skinned" decollement nature, similar in style to the Jura Mountains of Switzerland and France (Rich, 1934; Rodgers,

1949, 1953a, 1953b, 1963, 1964, 1972; Milici, 1963, 1970; Harris, 1970).

Compressive forces originating east of the basin have created a series of imbricate thrust faults which gradually decrease in intensity toward the west (Figure 3). "Hatcher (1972) states "some faults apparently originated as bedding thrusts in a weak stratigraphic layer and broke upward across competent strata from one shaley layer to another." The COCORP

(Consortium for Continental Reflection Profiling) seismic survey across the southern Appalachian region from Georgia to Tennessee has shown .. that the major tectonic feature of the area is an overthrust that transported crystaline rocks over sediments as much as 260 km horizontally (Cook et al. 1979). This late Paleozoic episode of deformation, crustal shortening, and mountain building coincides with

Wilson's (1966) proposed closing of the proto-Atlantic .

The southern Appalachian Plateau is an area of economic as well as academic geologic interest. Carboniferous coal deposits are extensive, and mining these coals is presently and has historically been the major 7

A A' CUMBERLAND RIDGE PLATEAU AND VALLEY COASTAL PLAIN

Carolina Slate belt (Lower Paleozoic)

100km.

GO mi.

Figure 3. Cross Section A-A' (Dott and Batten, 1971), from the Coastal Plain of South Carolina to the Cumberland Plateau of Kentucky (see Figure 2 for location). Pine Mountain (Figure 4, grid G9-J7) is the western most of the eastward dipping imbricate thrust slices which comprise the Appalachian Valley and Ridge geomorphic province. 8

industry of the region. Natural gas has been produced in eastern

Kentucky from the Devonian (black) Ohio Shale for nearly 100 years.

Some wells are still economically productive after more than 50 years

(Negus deWys, 1979).

Previous Works

There are several previous remote sensing studies of the southern/central Appalachian Plateau and Valley and Ridge provinces which have utilized radar images for geologic interpretations. Wing and others (1970), working with air-borne radar images of the Burning Springs

Anticline of West Virginia, found good correspondence between air photo lineament patterns, radar imagery lineament patterns, and surface joint strikes measured· in the field. Elder et al. (1974), using airborne radar images of the coal mining area in Buchanan County, Virginia, identified radar lineaments and lineament patterns as fault and joint systems; this was verified in the field by surface and in-mine investigations. Johnston

·et al. (1975) investigated the sites of several prominent lineaments-, observed on airborne radar, air photo and Landsat images of Virginia and

West Virginia. Field investigations of two prominent lineaments proved these to be the traces of previously unrecognized faults. Ford (1980), using Seasat-A radar images of the Valley and Ridge province of

Kentucky, Virginia, and Tennessee correlated Seasat-A radar lineaments

with mapped surface structures and geophysical trends in the basement.

Also, lithologic mapping was done by discriminating several image

textures which correspond to a dominant underlying lithology or lithologic --~ -~-~~- ~~- ~~---- ~-~-~

9

associations.

This thesis intends to extend to the Appalachian Plateau the type of study done by Ford (1980), and to serve as a follow-up to the work of

Johnston et al. (197 5). METHODS

To accomplish this project, an image data set was assembled

consisting of images acquired by the synthetic aperture radar (SAR)

carried onboard NASA's Seasat-A satellite, optical images acquired by the

multispectral scanner system (MSS) onboard NASA's Landsat series

satellites, photographic products from NASA's Skylab space station and

conventional high and low altitude photography. For this study, I interpretation maps of surface drainage patterns, surface texture

patterns, and surface lineaments were made from Seasat-A SAR images.

These data are analysed and compared to similar maps derived from

Landsat MSS images of the same approximate area, and with available

field-based geologic reports and maps. Two field trips were made to the

study area; Octoger 1-27, 1980; and April 1-30, 1981. The results of these

field investigations are incorporated into later sections of this report.

The primary imagery used for this study are Seasat-A SAR and

Landsat MSS images. They are presented in Figures 4, 5, 6, 7, and 8 A, B,

. and c. The imagery is in two sets which cover overlapping areas. Figures 4, 5, 6, and 7 are one set which covers approximately 10,000 km2. Figure

8 A, B, C comprise the second set which covers approximately 3,000 km2.

There is approximately 1,500 km2 of common territory for the two sets.

Figures 4 and 8A are SAR images acquired during revolution 874 of the

Seasat-A spacecraft and Figures 5 and 8B were acquired during revolution

694. These two orbital paths are depicted in Figure 1. Figures 6, 7, and

8C are Landsat MSS images covering the same approximate area as the

Seasat images. The scene in Figure 6 was acquired during the spring and

the scenes in Figures 7 and 8C were acquired in the winter. The winter

10 11

Figure 4. Digitally Pr:-_ocessed Seasat-A radar image from Revolution 874, (image name "Zebulon, Ky.") Radar Look direction N67.70W. This image covers portions of Kentucky, Virigina, and West Virginia. The local terrain is densely vegitated and maturely dissected with slopes that range mostly from 160 to 260. The major linear feature (grid G9-J7) is the crest of Pine Mountain the northern terminus of the Pine Mountain overthrust. Other major (and minor) linear features on this image can be correlated to ridges, stream segments, valleys, and alignments of these. The Levisa Fork of the Big Sandy River is the major drainage channel of the area which can be traced across the image (grid location Cl-D4-J7). Several fresh reservoirs are visible at (grid) 04, H6, and J8. The coarse surface texture seen to the east represents a larger ratio of resistant sandstone to more easily erroded shale and limestone. Conversely the finer texture seen to the west represents a higher ratio of limestone and shale to sandstone in the surface rocks. 12

0 10 20 RADAR ILLUMINATION km DIRECTION 13

Figure 5. Digitally processed Seasat-A radar image (image name "Salyersville, Ky."). Radar Look N67.70E., from Revolution 694-. The portion of the image displayed here is that which overlaps the area of Figure 4. See Figure 1 for location of overlapping area of revolutions 694 and 874. The area depicted here is wholly within the commonwealth of Kentucky. 14

_j A c D E F G H J L

RADAR ILLUMINATION DIRECTION

0 10 20

km 15

Figure 6. Landsat M~S band 7 (green, 0.8-1.1 micrometers) subscene acquired by Landsat 2 on December 2, 197 5. Eros Data Center ID /18231415292500. Solar elevation at the time of image acquisition (.±, 9:00 AM) was 24o, azimuth bearing N290W. The low solar elevation resulted in long surface shadows that highlight the surface topography. This image covers the same approximate area as Figure 4. (Seasat-A SAR, radar look direction N6 7. ?OW). 16

H

0 10 20 SUN ILLUMINATION km Dl RECTION 17

(

Figure 7. Landsat M~S band 7 subscene from image /18300671535XO (Eros Data Center ID II). This image was acquired by Landsat-3 on May 11, 1978. Solar elevation at the time of image acquisition (!: 9:00 AM) was 570, azimuth bearing N670W. The relatively high sun angle results in very little surface shadowing and the surface topography is subdued on the image. This figure covers the same approximate area as Figures 4 and 6. 18

/

0 10 20 SUN ILLUMINATION DIRECTION 19

Figure 8A. Mossaic of optically processed Seasat-A SAR images of a portion of the Pine Mountain thrust plate in Kentucky and Virginia (from revolution 874). Radar Look direction is N.67.70W, Radar Look angle is 200.

Figure 8B. Mossaic of optically processed Seasat-A SAR images of a portion of the Pine Mountain thrust plates in Kentucky and Virginia (from revolution 694). · Radar Look direction is N.67.70E, Radar Look angle is 200.

Figure 8C. Landsat MSS band 7 subscene of a portion of the Pine Mountain thrust plate in Kentucky and Virginia. At the time of image acquisition, the solar elevation was 240, azimuth bearing N.290W (from same scene as Figure 6, ID /18231415292500).

Figure 8D. Geologic structure interpretation map of the area covered by Figures 8A,B,&C. The linear ridge of Pine Mountain (the northwestern terminus of the Pine Mountain overthrust) passes through the area from southwest to northeast. Pine Mountain acts as a natural and geographic border between Kentucky to the north and Virginia to the south. Russel Fork fault marks the northeastern terminus of the Pine Mountain overthrust. Powell Valley anticline is a complex stucture which spans the length of the overthrust block, it is truncated on the south by the Hunter Valley thrust fault. Coburn fault and Indian Creek fault were formed by differential movement s>f the Pine Mountain overthrust block. Guest Mountain fault is newly reported, 30 meters of stratigraphic throw was observed in the field. 20

R A D A R I L L U M I N. A T I 0 N ILLUMINATION A c B 0 RADAR ILLUMI-

RUSSELL FORK FAULT

COBURN FAULT

CRANES~NEST FORK INDIAN CREEK -----+ .....,}"' FAULT 1 ...... , G U E S T M T. T H R U S T "~AU L T

POWELL VALLEY ANTICLINE;

0 10 20

KM 21

Landsat images most closely resemble the radar images because of solar shadowing at the low (240) sun elevation. The spring subscene has a much higher solar elevation (570). This image was used for drainage mapping only.

Radar Imagery and Seasat-A

Imaging radar systems are active remote sensing devices because they provide their own source of illumination; energy is transmitted to the surface and an image is created from the reflected energy received at the radar antenna (backscatter). The imaging geometry and time of acquisition ·are selectable. Imaging radar systems operate in the portion of the electromagnetic spectrum; and fog are transparent at the longer wavelengths. These combined properties render a radar imaging system independent of surface lighting and weather conditions.

The primary terrain factors which control radar image tone are surface -slope, surface roughness, and dielectric constant. Dielectric constant is primarily related to moisture content and is usually not a dominating factor.

Figure 9 illustrates the effects of topography and surface roughness on radar backscatter. The radar image formed thus records the radar reflectivity of the surface. Surfaces which reflect a lot of radar energy back to the receiving antenna (generally rough surfaces) are bright on a positive radar image. Surfaces which reflect back little radar energy

(generally smooth surfaces) are dark on the image. 22

Figure 9. Schematic representation of the interaction of several surface tyP-es with an imaging radar system. In A, the smooth (specular) surface (a road, a body of water) reflects most of the transmitted energy away from the radar antenna and a dark image tone will result. In B and C, rougher surfaces return more energy to the radar antenna and produce various gray tones on the radar image. Abrupt topographic discontinuities such as in D provide very strong reflections. Buildings, fault scarps, stream banks and the like can be subresolution in size, yet detectable because of their strong reflections. Radar look angle and radar depression angle are also pointed out in this figure. This figure was provided by M. L. Bryan of the Radar Remote Sensing Team at the California Institute of Technology, Jet Propulsion Laboratory. 23

I 24

Distortions due to imaging geometry are inherent in SAR imaging

systems. Surface slopes inclined towards the radar antenna appear

compressed relative to slopes that are tilted away. Extreme

foreshortening is known as radar layover. This occurs when the facing

slope is inclined at an angle greater than the look angle (Figure 9 A) of the

radar beam. Radar shadows are formed when the slope facing away from

the radar antenna is at an angle greater than the depression angle (Figure

I 9B). Discussions on imaging radar principles and technology are given in

Sabins, {1979, p. 177-231) and Jensen et al. {1977}. For several discussions

of Seasat-A SAR imaging geometry, see Ford et al. (1980, p. 3-5}.

Seasat-A, an oceanographic research satellite, was launched by the

National Aeronautics and Space Administration (NASA} on June 29, 1978.

The Seasat project was managed by the California Institute of

Technology, Jet Propulsion Laboratory (JPL}. The objective of Seasat-A

was a proof-of-concept demonstration of the capability to monitor the

ocean surface and near surface features such as surface waves, internal

waves, currents, ice cover, and other such ocean phenomena. Seasat-A

was equipped with a synthetic aperture radar {SAR} imaging system-­

among its payload of five sensors. The Seasat-A SAR provided the first

synoptic radar images of the surface of ~he obtained from an orbital

platform. Seasat-A operated successfully for 106 days until a massive

short circuit in the electrical system prematurely ended its mission. Prior

to failure, the SAR experiment had acquired a large amount of high

resolution (23.5 m} images of both ocean and land areas. The reader is

referred to Born et al. (1979) for a detailed description of the Seasat

mission. 25

The Seasat-A SAR operated at 23.5 em wavelength (L-band). It

acquired image swaths 100 km wide (range direction) and thousands of

kilometers long (azimuth direction). Seasat orbited at approximately 800

km altitude with an of 1080. The inclined orbit allowed

for two different illumination (range) directions. As the name implies,

Seasat was designed for oceanographic observations. The major

shortcoming of this from a geologists point of view was the fixed 200 look

/ angle (70° depression angle, see Figure 9) which resulted in considerable geometric distortion in topographically steep areas. Despite this, many

interesting geological studies have been made with Seasat-A radar images

(for example, Muskat et al. 1981; Blom and Elachi, 1981; Ford, 1980; and

Sabins et al. 1980). Elachi ( 1980) provides a general overview of the

Seasat-A mission and reports on several studies which utilized data

gathered by this ·satellite.

Seasat data are available from NOAA/EDIS, National Climatic

Center, Satellite Data Services Division, Room 100, World Weather

Building, Washington, DC 20233.

SIR-A (shuttle imaging radar) was carried onboard the space shuttll;,'!

Columbia during its second experimental flight in November, 1982. SIR-A

was designed similarly to the SAR on board Seasat-A. Some fundamental

differences were the 470 <.±. 30) look angle and 40 m image resolution of

SIR-A. Elachi (1982), provides a description of the SIR-A experiment and

some preliminary scientific results. 26

Landsat Series Satellites

On July 22, 1972, NASA launched the first Earth Resources

Technology Satellite (ERTS), later re-named Landsat. Four satellites in

the Landsat series have been launched to date. Each Landsat satellite

contains two optical imaging systems. Landsats 1, 2, and 3 have a Return

Beam Vidicon (RB V), and a Multispectral Scanner (MSS) onboard. Landsat

4 has a MSS system and a Thematic Mapper (TM) onboard. The Landsat

imagery used in this study is MSS data only.

The MSS is an optical mechanical scanner which images in four

adjacent spectral regions; 0.5 - 0.6 micrometers (green), 0.6 - 0.7

micrometers (red), 0.7 -0.8 micrometers (near ), and 0.8 - 1.1

micrometers (near infrared). Resolution of the MSS imagery is 79 m. The

MSS scans a sw-ath 185 km in the range direction. The data are

transmitted to earth in digital form from a satellite in a sun-synchronous

orbit at an altitude of 900 km.

Landsat 4 was launched July 16, 1982. Along with the multispectral

scanner, this satellite carries a second generation, improved earth-,

observing sensor called the Thematic Mapper (TM). The TM provides for

improved spatial resolution and spectral separation. A description of

Landsat 4, the spacecraft and its onboard sensors can be found in issue

number 23 of the Landsat Data Users Notes published by the U.S.

Geological Survey EROS Data Center 1 (1982).

lAs of January 31, 1983, the Landsat Data Users Notes is published by the U.S. Department of Commerce's National Oceanic and Atmospheric Administration (NOAA). ANALYSIS

Drainage

Major drainage of the study area is to the northwest. The major channels drawn from the radar image in Figure 4 are shown in Figure 10.

The Levisa Fork of the Big Sandy River occupies a deeply entrenched wide meandering valley which is fairly easy to trace through the rugged terrain

(grid Cl-D4-J7, Figure 4; grid C1-C3-D4-J7, Figure 6; grid C3-D4--J7,

Figure 7). The Levisa Fork basin and the Tug Fork basin are the principle drainage basins of the study area. Portions of the Kentucky River drainage basin to the west, and the Guyandotte River drainage basin to the northeast are within the study area.

Several fresh water reservoirs that are visible on the Seasat-A and

Landsat imagery are shown in Figure 10 (grid D4-, H6, J8, Figures 4, 6, and

7 respectively). On the Seasat-A SAR images, the lakes appear black because of the near total reflection of radar energy from the smooth surface of the water, resulting in little or no returned energy to the radar antenna (Figure 9A). On the Landsat images (Figures 6, 7, 8C) the lakes are black because water absorbs electromagnetic energy at the wavelengths these MSS images were acquired (0.8-1.1 micrometers, band

7).

Figure 11 was made by tracing blue lines on United States

Geological Survey 1 :250,000 series topographic maps. Four surface drainage interpretation maps were made for this study. Figure 12 was made from Seasat-A radar with a look direction of N67.70W (Figure 5).

Figure 13 was made from Seasat radar with a look direction of N67.70E

27 ~·····~·~·······~········-··~························~················~··········~~~·~~-··~.... ~ ...... ~... -·------~ ~ ~ ------~-- ... -...... ~.-..~.-- ...... - ....- ....· ... · ...- ...- .. - -- -~------

28

Figure 10. Generalized interpretation map of Seasat-A SAR image radar look direction N67.70W (Figure 4). This is the first interpret.~tion map made for this study. It highlights major drainage channels, lineaments, and surface textural boundaries. More detailed maps are presented in later sections of this thesis. 29

\ \ I I I \ I \ \ / ...... / ' , ..... / ' ...... · I I I r-­ \ I \ I ..... / I ' ' \ I

·. RS

/I I ....._ :· I \ \ ~- ...... ' • • • JI \ -, I \ \ ---'": I \ - ·"' I ...... ::- ....', ,...,.. :. .... 7 ,.- ·. 1

20 DRAINAGE LINEAMENTS ---- KM TEXTURAL BOUNDARY •• • •• • • • • • • • ------~------~------~------

30

I

Figure 11. Drainage channels as shown by blue lines on the U.S. Geological Survey 1 :250,000 series topographic maps covering the area ..shown in Figures 4, 6, and 7 (Bluefield, Virginia, NJ 17-8; Jenkins, Kentucky, NJ 17-7; Charleston, West Virginia, NJ 17-5; Huntington, West Viriginia, NJ 17-4). 31

I 32

Figure 12. Drainage interpretation map of Seasat-A SAR image, radar look direction N.67.70W (Figure 4). The original map was made at a scale of 1:250,000. Drainage channels are mapped as perceived on the image. Streams appear discontinuous or intermittent because of the relationship between stream channel orientation and radar look direction. Stream segments which trend parallel to the radar look direction are obscured on the image. Exceptions are very wide channels such as the Broad Bottom area of the Levisa Fork (Figure 4, grid F5), or those channels which were carrying a lot of water at the time of image acquisition resulting in a specular surface (see Figure 4, grid A5). The majority of stream segments mapped are oriented near perpendicular to the radar look direction. More stream channels were mapped from the western portion of the image because slopes are less steep there. In areas of very steep terrain (as in the eastern portion of the image), radar shadows, and radar layover mask and distort the stream channels on the image. 33

0 10 20 RADAR ILLUMINATION km DIRECTION 34

· Figure 13. Seasat drainage interpretation map of Seasat-A SAR, radar look direction N67. 70E (Figure 5). Drainage channels were mapped as perceived on the radar image resulting in a map of isolated ·segments. The majority of stream segments mapped strike perpendicular to the radar look direction. Original map was made at a sale of 1:250,000. 35

RADAR ILLUMINATION DIRECTION

0 10 20

km 36

Figure 14. Drainage interpretation map of Landsat MSS band 7 subscene, solar elevation 240, azimuth bearing N290W (Figure 6). Original map was made at a sacale of 1:250,000. Streams appear intermittent or discontinuous as on the Seasat-A SAR images for the same reason; the orientation of a stream segment relative to the illumination direction of the scene either highlights or suppresses its appearance on the image. The low winter sun elevation creates long shadows which highlight gently sloping topography. In the areas of rugged terrain (east) more drainage channels are perceived on the _MSS images than the SAR images, because the Landsat images are relatively distortion free whereas the Seasat images are distorted by radar layover. The majority of the stream channels shown on this map trend perpendicular to the solar illumination direction. 37

I

0 10 20 SUN ILLUMINATION DIRECTION km 38

I

Figure 15. Drainage interpretation map of Landsat MSS band 7 subscene, solar elevation 570, azimuth bearing N620W (Figure 7). Original map was made at a scale of 1 :250,000. Streams appear fntermittent or discontinuous as a result of the orientation of the various stream segments. The majority of the stream segments mapped trend perpendicular to the solar illumination direction. The related high solar angle of 570 results in negligible surface shadows in areas of low relief, hence less drainage channels are perceived on the image. To the east where the surface topography is of higher relief the c;Irainage channels are highlighted by surface shadows and are more easily perceived on the image. -~ ~ -- ... ~------~------~------~------~-~------~------······-=------=------·--·····-···------___ · ------

39

I

0 10 20 SUN ILLUMINATION DIRECTION km 40

(Figure 7). Figure 14 was made from a Landsat MSS subscene acquired during the winter of 1975 (Figure 8), and Figure 15 was made from a

Landsat subscene acquired during the spring of (1978) (Figure 9).

In Figure 11 regional dendritic drainage pattern is obvious. The drainage pattern appears more rectangular or trellis like in the four image interpretation maps (Figures 11-14) because drainage channel perception is strongly influenced by the direction of scene illumination. On all four drainage maps, it is obvious that the majority of drainage channel segments mapped are oriented approximately perpendicular to the image illumination direction. Drainage channels which strike near-parallel to the image illumination direction are virtually invisible, resulting in directionally biased interpretation maps.

Surface shadows are an important factor that influence drainage

channel perception on an image. Radar shadows are formed when the surface slope facing away from the radar antenna is at an angle greater than the radar depression angle. A low solar elevation angle will form

long surface shadows which enhance topographic features such as stream

channels and valleys. This is illustrated by comparing the two Landsat -,

interpretations (Figures 13 and 14). In the spring subscene (Figure 7), the

sun elevation is 570 and surface topography is suppressed. The winter

subscene (Figure 6) has a solar elevation of 240 which greatly enhances

the surface topography allowing for more complete drainage channel

mapping. Note too that Dewey Lake (Figures 4, 6, 7 grid D4) and Fishtrap

Lake (Figures 4, 6, 7 grid G6) are very prominent on the spring Landsat

subscene, but are obscured on the winter Landsat subscene by long

shadows, due to the lower sun elevation angle. 41

The higher resolution of the Seasat radar imagery allows for more detailed drainage mapping. Low order2 streams are discernible on the

Seasat radar imagery as opposed to stream orders of higher magnitude on the Landsat MSS images. The Seasat SAR is superior to Landsat MSS imagery for mapping drainage channels in areas of low relief only. In the high relief areas, radar layover and spatial distortions make drainage mapping difficult or impossible. Notice how Fishtrap Lake (Figure 5, grid

G6) is partially obscured by layover of the surrounding mountain tops.

The basic techniques for interpreting surface drainage channels from radar images are similar to standard aerial photography interpretations (Bryan, 1979). Image resolution, stream valley orientation relative to scene illumination direction, and scene illumination geometry are the major factors controlling the ability to map drainage channels.

The interpretation maps presented here (Figures 12-15) show how changes in these parameters will result in very different looking drainage interpretation maps of the same area. Several images with differing scene illumination directions and geometry are necessary for complete drainage-channel mapping of an area.

2This classification scheme is that of Horton (1945) where the smallest unbranched tributary is designated first order. 42

Surface Lithology and Image Texture

Sedimentary rocks of early Pennsylvanian age constitute the majority of bedrock exposures of the region. Only along the north face of

Pine Mountain are older rocks of Devonian age and Mississippian age exposed (Figure 16). The lower Pennsylvanian section is comprised of sandstone, conglomerate, shale, siltstone, coal, underclay, and thin limestone (see Figure 17).

The rocks are largely deltaic in origin (Rice, et al. 1979). The lithology of a single bed commonly varies from place to place even within short distances. Lateral and vertical gradation is common with sandstone grading to shale and visa versa.

Campbell (1893) first described the lower Pennsylvanian section exposed south of Pine Mountain and named the Lee Formation, Norton

Formation, Gladeville Sandstone, and the Wise Formation. North of Pine

Mountain in eastern Kentucky these lower Pennsylvanian rocks are divided into the Lee Formation and the Breathitt Group (Rice, et al. 1979). In southwestern West Virginia, the lower Pennsylvanian sedimentary section is known as the Pottsville Group (Cardwell, et al. 1968). A correlation chart is presented in Figure 18.

Few rock outcrops can be seen on the Seasat radar and Landsat MSS imagery because the study area is heavily forested and covered by a thick regolith mantle. Rock outcrops are limited to stream valleys and road cuts. Although the bedrock is almost totally masked on the imagery and not seen directly, certain textural patterns on the images are evident which represent characteristic lithologies. Gross surface textural pattern 43

.Figure 16.. Tilted interbeds of limestone and shale (Mississippian) exposed along the crest of Pine Mountain (trace of the Pine·, Mountain Thrust Fault). 44

I

Figure 17. Typical Appalachian Plateau stratigraphic section as seen in a roadcut along U.S. 23 near Pikeville, Kentucky. This is the Bingham coal zone in the Lower Pennsylvanian Breathhitt Fm. This roadcut exposes 100' of stratigraphic section. The lower 50' is the Bingham coal zone which consists of coal beds 4' to 15' thick, interbedded shale, siltstone, and sandy claystone. Seen above the Bingham coal zone is a (.±,50') laminated siltstone and claystone unit. The Bingham coal zone, along with other major coal zones of the Breathhitt Formation, are mined extensively in this area. See Figure 18 for regional stratigraphic section. l~5

I 46

GEOlOGIC' EASTERN VIRGINIA WEST VIRGIN lA SYSTEM KENTUCKY z WISE FM. sz BREATHITT LLI~ GLADEVILLE POTTSVILLE t-:.J GROUP! <(>- ss ..Jen GROUP z NORTON FM. z LLI LEE FORMATION LEE FM. ' 0.. t BLUESTONE FM. PENNINGTON PRINCETON SS MAUCH CHUNK z u HINTON FM. <( GROUP Bl IE FIELD FM. FM. -Q.. 0. 't NEWMAN GREENBRIER GREENBRIER -en M LIMESTONE .FM. -GROUP . en -.../.,- -en en ' t BORDEN FM. MAC CRADY FM. MAC CRADY FM. - SUNBURY SHALE . :E L' BEREA SS. PRICE FM. POCONO FM. l BEDFORD SHALE -- HAMPSHIRE FM. z OHIO CHEMUNG <( FOREKNOBS FM. ·z- SHALE FM. SCHERR FM. 0 > PARKHEAD SS LLJ 0 BRALLIER FM. . . - ..

After Figure 18. Geologic correlation chart of the study area. Cardwell, et al (1968). 47

maps were drawn from the two Seasat images and the winter Landsat subscene (Figures 4, 5, and 6). These maps are presented in Figures 19, 20 and 21. Examination of the geologic maps of the area (Appendix) has led to the observation that the coarse texture seen in the eastern portion of the study area is due to a higher percentage of resistant sandstone relative to more easily eroded shale. Conversely, the finer texture seen in the western portion of the study area is due to a higher ratio of carbonates and shale to sandstone in the surface and near surface bedrock.

The gray bands seen on the radar image along Pine Mountain (Figure

4, grid A4 to B2) are caused by varying slope angles corresponding to the different sedimentary layers exposed along the mountain front. To the north of Pine Mountain, several dark patches are noticeable on the radar image east of Dewey Lake (Figure 4, grid D4). This is where active strip mining has altered the landscape to a relatively smooth surface which scatters the radar energy in the same manner as the smooth surfaces of the several large bodies of water (Figure 9 A).

Structure

Major faults of the area are expressed topographically in the form of extensive linear and curvilinear features. The most striking of these is the linear ridge of Pine Mountain (Figure 4, grid F9-J7). At the northern margin of Pine Mountain is the trace of the Pine Mountain overthrust

(Figure 22). This is the northern terminus of the Pine Mountain thrust plate (also known as the Cumberland overthrust block). The thrust plate Figure 19. Surface texture interpretation map from Seasat-A SAR, radar look N67.70W (Figure 4}. The line represents a boundary between a coarse surface texture to the east (B) and a finer surface texture to the west (A}. The coarse texture results a higher ratio of sandstone to shale in the surface and near surface rocks and conversely the finer texture results from a higher ratio of limestone and shale to sandstone in the surface and near surface rocks. All surface rock is Pennsylvanian in age, belonging to the Breathitt Formation. 49

B A

0 1 0 2 0 R A D A R .\ K M .ILLUMINATION 50

Figure 20. Surface texture interpretation map from Seasat-A SAR, radar look N67.70E (Figure 5). The line represents a boundary between a coarse surface texture to the east (B) and a finer surface texture to the west (A). The boundary is very similar to that in Figure 18. The slight difference can be attributed to the different look directions of the two radar images, ··giving the two images slightly different textural appearances. All surface rock is Pennsylvanian in age, belonging to the Breathitt Formation. The finer surface texture in the west is attributed to the higher percentage of easily eroded limestone and shale to resistant sandstone. The coarse texture (B) is due to a higher percentage of sandstone in the surface and near surface rock. 51

i

A

ILLUMINATION DIRECTION

0 1 0 2 0

K M 52

/

Figure 21. Surface texture interpretation map from Landsat MSS band 7 winter subscene (Figure 6). The line represents a boundary between a coarse surface texture to the east (B) and a finer surface texture to the west (A). The boundary is very similar to that in Figure 19. Again, the slight difference in the boundary line can be attributed to the differing illumination directions of the images. All surface rock is Pennsylvanian in age, belonging to the Breathitt Formation. Sandstone is the predominant surface rock in the eastern portion of the scene, and limestone and shale are the predominant lithologies to the east. 53

A B

0 1 0 2 0

K M N\ ILLUMINATION""SUN 54

Figure 22. Schematic Map of Major Geologic Structures. This map covers the northern image set (Figures 4,5,6,7). Rome Trough, a grabben in the Precambrian basement, is the major basement structure in the study area. The southern border as mapped by Ammerman and Keller (1979) passes through the study area from the southwest to the northeast. R. C. Schumaker (personal comm, 1980) of West Virginia University associates the Warfield fault of West Virginia with rifting along the southern margin of Rome Trough. Anticlinal and synclinal axis shown are after McLoughlin (1979). The Pine Mountain overthrust is the northern terminus of the Cumberland overthrust block, bounded on the northeast by the Russell Fork fault. The Coburn fault was confirmed by Johnston et al. 1979. The Indian Creek fault is newly ~onfirmed and named in this report. Surface rock of this area consists of sandstone, shale, limestone, coal, siltstone, and claystone of the Pennsylvanian Breathitt Formation. Along the ridge of Pine Mountain older sandstone and shale of Mississippian and Devonian age are exposed. 55

WARFIELD FAULT

PINE MOUNTAIN

INDIAN CREEK FAULT

0 10 20

KM 56

is the western most of the imbricate thrust sheets which comprise the

Valley and Ridge province to the east (Hatcher, 1972). The Pine Mountain overthrust is thought by most workers to be the decollement or sole fault which extends beneath the Valley and Ridge (see Figure 3). The Pine

Mountain thrust plate is 200 km long by 40 km wide. It is bounded on the east by the Hunter Valley fault, on the southwest by the Jacksboro fault

(southwest of the study area), and on the northeast by the Russell Fork fault. Wentworth (1921) first described this overthrust block. Subsequent work by Butts (1927) and Rich (1934) developed into the funaamental concepts of the mechanics of "thin-skinned" tectonics, a term later coined by Rodgers (1949).

Deep exploration wells drilled into the Pine Mountain thrust sheet have revealed bedding plane thrust faulting in three thick shale units from

Cambrian (Rome FM.) to late Devonian (Chattanooga/Ohio Shale) in age.

The sole fault ramps upward along diagonal shear planes through competent beds and continues as a bedding plane thrust in higher incompetent layers (Harris, 1970). Englund (1961, 1971), through detailed mapping- across the Pine Mountain overthrust plate has determined that the thrust plate experienced rotational movement resulting in displacement of 6 km along the Russell Fork fault and 18 km along the

Jacksboro fault. There are several faults within this thrust sheet which strike perpendicular to the long axis of the allochthon. They are the

Rocky Face fault (out of the study area), the Coburn fault (Johnston, et al. 1975), and the recently confirmed Indian Creek fault (see page 17).

These faults are the result of differential motion of the thrust plate during its emplacement. 57

Powell Mountain (Figure 8A, B, C, D) is the nose of the plunging

Powell Valley anticline, a complex structure which spans the length of the allochthon. The Powell Valley anticline is thought to have been formed during the final stages of movement of the allochthon, by a doubling up of the stratigraphic section along upward ramping thrust faults which do not break the surface (Harris, 1970). The Hunter Valley thrust fault truncates

Powell Mountain at its eastern slope.

North of the Pine Mountain thrust plate on the Appalachian Plateau, the superficial structures are more subtle. The Warfield fault (Figure 22), a normal fault shown on the geologic map of the state of West Virginia

(Cardwell, et al. 1968), has been mapped in the subsurface only; on the

Seasat and Landsat imagery, several lineaments with similar strike are observed at the approximate location of this fault in West Virginia and eastward into Kentucky (Figure 4 gird E3-G 1).

Rome trough, a grabben in the Precambrian basement, is the major· basement structure in the study area. An extensive study of gravity survey data by Ammerman and Keller (1979) has delineated the boundaries . - of this major structure. The southern border passes through the study·, area from the southwest to the northeast (Figure 22). Thickening of the

Rome Formation (Devonian) over this structure suggests it was still developing during Paleozoic times. An east/west alignment of several small streams occurs in the western portion of the study area (Figures 4,

5, 6, 7, all grid D5) -- this is perhaps a surface expression of this major basement structure. 58

Indian Creek Fault

Substantial evidence of surface faulting has been found along the

trace of a major geomorphic lineament located on the Pine Mountain

thrust plate. The lineament is formed by an alignment of several stream

segments, dry valleys and intervening gaps. The structural features

observed in outcrop which suggest the presence of a fault include highly

I fractured and jointed rock, highly weathered rock, and tilted sedimentary

beds (Figure 23) at various locations along the trace of the lineament.

This lineament is herein named the Indian Creek Fault (Figure 80).

The Indian Creek drainage basin was first described by Eby (1923)

who explained the remarkable linearity of Indian Creek as a structure

controlled phenomena caused by the presence of a sharply flexing

monocline between an anticline to the east and a syncline to the west.

Rich (1934) suggested that these structures were formed in the same­

manner, and during the same event which formed the Powell Valley

anticline. Johnston et al. (197 5) noted the Indian Creek lineament as

prominently visible on air photos and ER TS (Landsat) imagery (Figure .,

8C). They suggested this might be the trace of a previously unrecognized

fault and recommended field checking to confirm this. No published

subsurface geophysical data were found for this area of Virginia.

The Indian Creek lineament is also very prominent on the Seasat-A

SAR imagery (Figure 8 A, B). For this project, a field trip was undertaken

to the site of this lineament to look for evidence of surface faulting. The

evidence found to confirm the Indian Creek fault is faulted rock exposed

in a road cut along U.S. 23, on the west side of the road due west of the 59

/

Figure 23. Tilted shale beds of the Wise Formation (Pennsylvanian) exposed in a road cut along the trace of the Indian Creek fault. Note the fractured, jointed, and weathered nature of the rock. 60

town of Wise, Virginia. At this location, sandstone of the Gladeville

Formation is exposed on the east side of the road, and sandstone and shale of the Wise Formatiion are exposed on the west side of the road. The lithologic contacts are shown on the Virginia state geologic map (Milici et al. 1963). Within sandstone of the Wise formation (west side of the road), two distinct orientations of slickensides are apparent (Figures 24-, 25 and

26). One pattern on a vertical fault trends due north and plunges variably from 900 to /.t.OOE. The second group is a horizontal pattern which trends

N70W and plunges 580W. From these observations only the relative sense of motion could be determined -- the vertical throw and horizontal displacement are probably limited to only a few tens of meters at most.

Given the trend of the vertical slickensides and the fact that the

Gladeville sandstone is older than the Wise Formation (Campbell, 1893) then the east side must have moved up relative to the west side making this a reverse fault. More detailed field mapping is necessary to determine the amount of vertical and strike slip.

Guest Montain Fault

Along Virginia State Road 6lt-6, 10 km east of the Indian Creek fault, offset beds are exposed in the highway road cut. A coal bed and a sandstone bed are separated in a reverse sense by approximately 30 m

(Figures 27, 28). Figure 29 shows a gouge zone (about 15 em wide). There is a 2-mm thick layer of silt along the fault plane. The fault strikes

N810E and dips /.t./.t.ONW. This is a reverse separation fault with the footwall on the south side. The fault could not be traced for any distance 61

Figure 24. Exposure of slickensided sandstone in the lower Pennsylvanian Wise formation seen in a roadcut along U.S. 23 near Wise, Virginia. The sandstone exposed in this roadcut is deformed and altered along the plane of the slickensides, which are obvious fault related features. Exposed slickensided surfaces in this roadcut confirm the existence of the Indian Creek fault. 62 63

Figure 25. Slickensides in deformed sandstone of the Wise Formation exposed in a roadcut along the trace of Indian Creek fault. This photograph is from the same roadcut as Figure 24-.

------··----~--~--- -~------64 65

I

Figure 26. Polished sandstone of the Wise Formation with slickensides orientated in a strike-slipmode. Note the i"eflection of the coin off the polished surface. Photo taken at same outcrop shown in Figure 25. 66

I 67

Figure 27. Guest Mountain fault, as seen in a roadcut along Virginia State Road 646, 10 km east of the trace of the Indian Creek fault. This fault is recognized by two offset beds seen in the center of this photograph (see the interpretive diagram presented in Figure 28).

------~- - 68 69

Figure 28. Line drawing made as an overlay of the photograph of the trace of the Guest Mountain fault (Figure 27). A bed of fine grained sandstone (20 em - 30 em thick) is over lain by a thin coal bed (lO em - 15 em thick) with laminated shale beds above and below; these two beds together are offset by 30 m of stratigraphic throw (1 0 m net vertical offset). This is a reverse separation fault with the hanging wall to the south. 70

J: .4 1- .. a:: • ,,.. 0 z ·:· ' . t • ..... '·' I r 't '. ' '

~ 0: •• ~ .' ...' .. ··. / 0\J • ., , / ' ,. , .. t" (\f a.. II 0 ... = I- LLJ ..J

{ I J: , I­ ,, ::> 0 I ·' (/) 71

)

Figure 29. Gouge Zone of the Guest Mountain fault. The coin shown~for scale is a quarter. Note the thin (2 mm) film of silt along the fault plane. Matrix material is weathered shale footwall abuted against fine grained sandstone (hanging wall). 72

I -----~~~----·--~

73

in either direction, nor is there an obvious lineament on any of the Seasat­

A or Landsat imagery which covers this area (Figure 8A, B, C). The strike

of the fault as measured at this exposure is parallel to the strike of the

north flank of the Powell Mountain anticline (Figure 8D). This is a

previously unrecognized fault, herein named the Guest Mountain fault

after the mountain upon which this exposure occurs.

I Other Possible Faults

The sites of several other prominent Seasat-A SAR/Landsat MSS

lineaments in Kentucky and Virginia were investigated for evidence of

surface faulting. Grapevine Creek (Figure 10) and Ferrell Creek (Figure

10) are linear flowing streams which strike parallel to the trend of the

Pine Mountain overthrust. Toller Creek (Figure 10) and Cranesnest Fork

(Figure 8d) strike perpendicular to the Pine Mountain overthrust.

Grapevine Creek, and Ferrell Creek trend similar to the subsurface

contours of the Ohio Shale (Figure 30). Grapevine Creek, Ferrell Creek

and Toller Creek parallel basement trends shown on the residual total

aeromagnetic intensity map of Kentucky (Figure 31) and the simple

bouguer gravity map of eastern Kentucky (Figure 32).

At least two field days were spent at each of these locations.

Although fractured, jointed, and weathered rock crops out along these

lineaments, no apparent offset beds or slickenside surfaces were observed

which could confirm the presence of faulting. 74

Figure 30. Structure contours drawn on the top of the Ohio Shale (after Fulton, 1979). Grapevine Creek and Ferrell Creek are both identified as prominent lineaments on the Landsat MSS and Seasat-A. SAR images. These two lineaments trend near parallel to the subsurface contours of the Ohio Shale. (ContoUI;'s :iJl.;feet.) 75

01 _____, 10_:c= 201 1--.::J.-~K~MI'~- I I I 76

Figure 31. Residual total intensity aeromagnetic map (after Johnson et al. 1980), contour interval is 100 gammas. Grapevine Creek, Ferrell Creek, and Toller Creek are prominent lineaments identified on the Seasat-A SAR and Landsat MSS images. These three lineaments parallel the basement trends shown on this f!lap. 77

.. ~. ' 78

I

Figure 32. Simple bouguer gravity map (after Ammerman and Keller, 1979). Bouguer density is 2.67 g/cc. Contour interval is 5 milligals. Grapevine Creek, Ferrell Creek, and Toller Creek are prominent lineaments identified on the Seasat-A SAR and Landsat MSS images. These three lineaments parallel the basement trends shown on this map. 79

/

0 KM 20 I 80

Lineaments

Lineament is defined by O'Leary, et al., (1976) as "a mapable linear or

curvilinear feature of a surface". Detailed lineament interpretation maps

were made from two Seasat-A SAR images and a Landsat MSS subscene of

the same approximate area. These maps are presented in Figures 33, 34,

and 35. These maps were then digitized in order to create the rose

I diagrams shown in Figure 36, A and B. Lineaments mapped here are

geomorphic features such as stream segments, valleys, ridges and

alignments of these. Lineaments are perceived on the images as tonal and

textural contrasts resulting from differential reflection of the radar

energy or sunlight from varying surface slopes.

At first glance, the Seasat SAR and Landsat MSS lineament

interpretation maps look very similar, but a closer examination reveals

subtle and significant differences primarily due to the scene illumination

directions. The solar azimuth bearing N290W is significantly different

from the two radar look directions of N67.70W and N67.70E. Geomorphic

lineaments which are parallel to the Illumination direction are supressed ·,

on the images whereas geomorphic lineaments which trend perpendicular

to the illumination direction are enhanced on the images. This results in

directionally biased interpretation maps. The rose diagrams in Figures 36

which depict lineament frequency per azimuth class, show this most

dramatically. In all cases, lineaments oriented perpendicular to the

illumination direction were mapped five times as frequently as lineaments

oriented along the illumination direction. 81

Figure 33. Lineament Interpretation Map from Seasat Radar Image, Radar Look N67.70W (Figure 4). Original map was made as a clear acetate overlay of a photographic print at a scale of 1:250,000. Lineaments mapped are geomorphic features such as stream valleys, dry valleys, and mountain ridges. This map contains 1322 lineaments. The majority of lineaments mapped -strike near perpindicular to the Radar look direction (see rose diagram A in Figure 36). 82

0 10 20 RADAR ILLUMINATION DIRECTION km 83

Figure 34. Lineament Interpretation Map from Seasat Radar Image. Radar Look N67.70E (image name "Salyersville, Ky.") These mapped lineaments are geomorphic features. The majority strike near perpendicular to the radar Look direction. 84-

RADAR ILLUMINATION DIRECTiON

0 10 20

km 85

I

Figure 35. Lineament interpretation map from Landsat MSS band 7 subscene, solar elevation 240, azimuth bearing N290W. The majority of lineaments mapped trend near perpendicular to the illumination direction. Almost no lineaments were mapped trending parallel to the illumination direction. This is depicted with rose diagram B in Figure 36. 86

I

0 10 20 SUN ILLUMINATION km DIRECTION 87

/ A RADAR llLUUIHATION OIRECTJON

SUN:ILLUMINATION DIRECTION ,·-N2~·w

Figure 36. These rose diagrams depict the number of lineaments mapped per azimuth class. Diagram "A" is from Seasat-A SAR image radar look direction N670W ("Zebulon"). Diagram "B" is from Landsat MSS band 7 subscene with a solar elevation of 240, azimuth bearing N290W. This figure is meant to qualitatively show that lineament perception on an image is strongly influenced by the orientation of the lineament relative to the illumination .direction. The majority of the lineaments mapped on both images trend near perpendicular to the illumination direction and almost no lineaments were mapped trending parallel to the illumination direction. 88

There were 1322 lineaments mapped from the Seasat SAR looking

N67.70W. There were 1240 lineament mapped from the Landsat MSS subscene of the same relative area. The lineament orientation trends are similar except for the obvious difference due to illumination direction; also, the effect of foreshortening and layover result in slighly rotated orientations on the radar images.

The effect of illumination geometry on lineament perception is evident when comparing the Landsat MSS winter subscene to the spring subscene (Figures 6 and 7 respectively). The winter solar elevation angle causes long shadows on the surface which highlight the topography. The spring solar elevation angle of 570 has a negligable shadowing effect on the surface, and the apparent image topography is subdued.

Figures 37 and 38 · were drawn to compare two lineament interpretation maps from different sources (Figures 33 and 35). The original maps were redrawn in colored ink then co-registered and compared to each other. Figure 37 shows geomorphic lineaments visible on the Seasat image looking N67.70W (Figure 33) which are not on the

Landsat MSS interpretation map (Figure 35). Figure 38 shows lineaments· mapped from the Landsat MSS image and not seen on the Seasat-A SAR.

All the individual lineaments are drainage channels, dry valleys, or ridges.

There was no significant difference between the type of lineaments mapped on the different images. These comparisons show that the dominant factors which govern lineament perception on an image are the lineament orientation relative to illumination direction, and the image illumination geometry.

Lineament trends have been shown to reflect faulting, jointing, and 89

Figure 37. Lineaments mapped from Seasat-A SAR with Look Direction N67.70W which do not appear on Landsat MSS Band 7 Winter Scene. The majority of these lineaments strike near perpendicular to the radar look direction and near parallel to the solar illumination of the Landsat MSS image. This shows how lineament orientation relative to illumination direction effects lineament perception on an image. 90

I I I I ;I' I \ \ ~ /1\ l ft I f I ift 1// I\..(! f ~ )'.l \\ \ I 1/ ~ \\ \ I f\ 1 I \ I I II I ,,, I ~ \ ''\)\ \\ /y '\ { \J I!\ \ I ( I 1/ 'Y; \ I II \ \ \ \ I \ '\ I I \ I "-1 j \ \ I 1 I \ \ I I I\ ~\I/ \ ,,\ I \ I; \ \\ " l \ I I, \f\ III I \ I I \ I \ I \ \ 'l \ // I I -~ I I I I\ !' I I I' I I I I \1 I I /I /1''\'i \ I/;/ I\ V/1/, I \I, ,,,, \ \\ ,. I I l 1 ,, \ / \ II - I I \ \ \ \ II: ~ I I I I ,, II I :I \ ;' I I I /;.. I ,, 1/ I I !' \ I I I \ \ \ II I I ;' I /\ I I ;I I I \'\ I \ If I "'\ \ tl /t 11 I

I I I, I ';\ \'I ''J /I I \ \ ,, I \' I ~ I l j I \ 1\ I; 11 \ \ I I \') I\\\\ I \I I \ \ \ '""- \ I II \ \ rl 1 1 1 \\I 1 \' \ ) I \ \ \ I \ \ \ I 1

RADAR ILLUMINATION DIRECTION \N km 91

Figure 38. Lineaments mapped from Landsat MSS winter subscene, which do not appear on Seasat-A SAR radar look N67.7°W. The results are similar to that of Figure 37. The lineaments shown here trend near perpendicular to the solar illumination direction and are prominent on the Landsat MSS image whereas they are not perceived on the Seasat-A SAR image because they trend near parallel to the radar look direction. 92

/j/ / / / I - --/

..-- 1 - y/./ / - _,.. ;J /-:_ /- - / -- - I // / .,

"';i I ------; /' ...--: r- - /

20 SUN "' 1LLUM1NAT10N ""'DIRECTION 93

fracture patterns in the sedimentary cover of the area (Wing et al. 1970;

Elder et al. 1974; Johnston et al. 197 5). Natural gas is produced in this

region from fracture-controlled reservoirs (Ryan, 1974; Negus deWys,

1979, Pryor et al. 1981). G. Owens (Personal comm.) reports that

Columbia Gas Company has good success in achieving higher yield wells,

without the added cost of secondary formation fracturing, by drilling near

geomorphic lineaments mapped from airborne radar images of Columbia's

I Haysi gas field located in Virginia off the northeastern border of the Pine

Mountain Overthrust (Figure 5, grid J7). Howard, et al. (1979) reports

similar results from gas wells drilled in Perry County, Kentucky.

Awareness of possible structural lineaments can also aid in the siting

of subsurface coal mines by recognizing and averting potential hazardous

conditions. Proximity of a "tUnnel to a structural lineament may make for

unstable roof rock conditions, or allow flooding of the shaft by migrating

ground water or toxic gases. ------

SUMMARY

This study shows that satellite borne radar images are a viable tool for

synoptic geologic reconnaissance mapping in areas of rugged terrain and

dense vegetation cover.

Interpretation maps of surface draingage channels made from Seasat­

A SAR images were compared to a similar map made from a Landsat MSS

image. In areas of low relief, high image resolution and low scene

illumination angle provide for greater discrimination of smaller tributary

channels on the Seast-A SAR imagery than on the Landsat MSS image.

This is due to the high radar sensitivity to slope change at low radar

illumination angles. However, in areas of high relief inherent geometric

distortions provide for slighly rotated orientations of the stream channels

on the radar image.

Gross surface texture maps were made from Seasat-A SAR images.

Comparison with U.S.G.S. 1:24,000 geologic maps of :the area suggest that

generalized surface lithologies can be related to gross image textural

patterns. The coarse surfac-e/image texture represents a higher

percentage of resistant sandstone to more easily eroded shale, and the

finer texture represents a higher percentage of limestone and shale to

sandstone in the surface and near surface rock.

Seasat-A SAR lineament interpretation maps were made and compared

with similarly derived Landsat MSS maps. In regions of low topographic

relief, the Seasat-A SAR images reveal more geomorphic lineaments

because of the higher image resolution and the lower illumination angle.

Topographic enhancement on Landsat MSS images is due to solar

94 95

shadowing which is most pronounced at sun elevation angles of less than

30o. In areas of high topographic relief the Seasat-A SAR images are distorted because of the radar layover effect, but the Seasat-A SAR images are still as effective as the Landsat MSS images for lineament detection.

The lineaments mapped on the Seasat-A SAR and the Landsat MSS images are geomorphic features such as streams, valleys, ridges and combinations of these which represent joint sets, fracture patterns or faults. Several prominent lineaments correspond to major structural, gravity, and magnetic trends in the basement rock. Field checking of the

Indian Creek lineament in southwestern Virginia has proven this to be the trace of a previously unreported fault, to be known now as the Indian

Creek faultc A previously unmapped thrust fault was also recognized in the field. This fault shows 30 m of stratigraphic throw, it is now known as the Guest Mountain fault. REFERENCES CITED

Ammerman, M. L., and Keller, G. R., 1979, Delineation of Rome Trough in Eastern Kentucky by Gravity and Deep Drilling Data: The Am. Assoc. Petroleum Geologists Bull., v. 63, No. 3, p. 341-353.

Blom, R. and Elachi, C., 1981, Spaceborne and Airborne Radar Observations of Sand Dunes; Jour. Geophys. Res. v. 86, p. 3061-3073.

Born, G. H., Dunne, J. and Lame, D., 1979, Seasat Mission Overview, Science, v. 204, p. 1405.

Bryan, M. L., 1979, The use of radar imagery for surface water investigations, in: Deutch, M., D. R. Wiesnet, and A. Rango (eds.), Satellite Hydrology, Minneapolis, MN: American Water Resources Association, 1981, p. 238-251.

Butts, Charles, 1927, Fensters in the Cumberland overthrust block in southwestern Virginia Geol. Survey Bull. 28, 12 p.

Campbell, Marius R., 1893, Geology of the Big Stone gap coal field of Virginia/Kentucky: U.S. Geol. Survey Bull. Ill.

Cardwell, D. H. Erwin, R. B. and Woodward, H. P., 1968, Geologic map of West Virginia: Virginia Geological and Economic Survey.

Dennison, J. M., 1976, Gravity tectonic removal of cover or Blue Ridge anticlinorium to form Valley and Ridge province: Geological Society of America Bull., v. 87, p. 1470-1476.

Dott, R. H., and Batten, R. L., 1971, Evolution of the Earth, New York: McGraw-Hill Book Company.

Eby, J. B., 1923, The geology and mineral resources of Wise County and the coal-bearing portion of Scott County, Virginia: Virginia Geol. Survey Bull. 24, 617 p.

Elachi, C., 1980, : Geol. and Oceanographic Applications; Science, v. 209, p. 1073-1082.

1982, Shuttle Imaging Radar Experiment: Science, Vol. 218, p. 996-1003.

Englund, K. J., 1961, Rotational block of the Cumberland overthrust sheet in southeastern Kentucky and northeastern Tennessee: U.S. Geol. Survey Prof. Paper 424-Cd, p. C74-C76.

1971, Displacement of the Pocahontas Formation by the Russell Fork Fault, southwest Virginia: U.S. Geol. Survey Prof. Paper 7 50 B, p. Bl3-B16.

96 97

Ford, J. P., 1980, Seasat Orbital Radar Images for Geologic Mapping: Tennessee-Kentucky-Virginia: A.A.P.G. Bull., v. 64, No. 4, p. 2064- 2094.

Ford, J. P., Blom, R. G., Bryan, M. L., Daily, M. I., Dixon, T. H., Elachi, C. and Xenos, E. C., 1980, Seasat Views North America, the Caribbean, and Western Europe with Imaging Radar: National Aeronautics and Space Administration, JPL Publication 80-67.

Fulton, L. P., 1979, Structure and Isopach Map of the New Albany­ Chattanooga-Ohio Shale (Devonian Mississippian) in Kentucky: Eastern Sheet: Kentucky Geological Survey.

Gwinn, V. E., 1964, Thin-skinned tectonics in the Plateau and northwestern Valley and Ridge provinces of the Central Appalachians: Geol. Soc. America Bull., v. 75, no. 9, p. 863-900.

Harris, L. D., 1970, Details of Thin-skinned tectonics in parts of Valley and Ridge and Cumberland Plateau provinces of the Southern Appalachians in Studies of Appalachian Geology: Central and Southern: Interscience Publishers.

Horton, R. E.. ,, 1945t Erosional Development of Streams and Their Drainage Basins: Hydrophysical Approach to Quantitative Morphology, Geological Society America Bull., Vol. 56, no. 1, p. 275-370.

Howard, J. F., Lahoda, S. J., Zirk, W. E., and Komar, c. A., 1979, Gas Production of Devonian Shale Wells Relative to Photo Lineament Locations: A statistical analysis: Morgantown Energy Research Center, METC/CR-79/28.

Jensen, H., Graham, L. L., Procello, L J., and Leith, E. N., 1977, Side­ Looking Airborne Radar: Scientific American, v. 237, no. 4, p. 84-95.

Johnson, R. W., Haygood, c., and Kunselmen, P.M., 1980, Residual Total Intensity Aeromagnetic Map of Kentucky: Eastern Sheet: Kentucky Geological Survey.

Johnston, J. E., Miller, R. L., and Englund, K. J., 197 5, Applications of Remote Sensing to Structural Interpretations in the Southern Appalachians: U.S. Geol. Survey Jour. Research, v. 3, p. 285-293.

McLoughlin, T. F., 1980, The Geological Significance of Landsat Imagery Lineament Analysis in Selected Areas of Eastern Kentucky": Masters Thesis, Morehead State University, Morehead, Kentucky.

Milici, R. c., Spiker, C. T., and Wilson, J. M., 1963, Geologic map of Virginia: Commonwealth of Virginia Division of Mineral Resources. Milici, R. c., 1975, Structural Patterns in the Southern Appalachians: Evidence for a Gravity Slide Mechanism for Alleghanian 98

Deformation: Geological Society Qf American Bull., v.. .86, p. l316- 1320.

Miller, R. L., 1973, Where and why of Pine Mountain and other major fault planes Virginia, Kentucky, and Tennessee: American Jour. of Science, Cooper, vo1. 273-A, p. 353-371.

Muskat, J., Ciancanelli, E., Blom, R., 1981, Seasat Radar Images for Mapping in Geothermal Areas: Geothermal Resources Council, Transactions, v. 5, p. 115-118.

Negus DeWys, J., 1979, The eastern Kentucky gas field: West Virginia Univ:ersity Department of Geology and Geography.

Negus DeWys, J., and Shumaker, R. C., 1979, Relations of Gas Occurrence

I to Geologic Parameters in Eastern Kentucky Gas Fields (abs.): Am. Assoc. Petroleum Geologists Bull., v. 63, no. 3, p. 502.

O'Leary, D. W., Friedman, J. D., Pohn, H. A., 1976, Lineament, Linear, Lineation: Some proposed new standards for old terms: The Geological Society of America Bull,. v. 87, p. ll/.63-14-69.

Pryor, W. A., Maynard, J. B., Potter, P. E., Kepferle, R. c., Kiefer, J., 1981, Energy Resources of Devonian-Mississippian shales of eastern Kentucky, Lexington: Kentucky Geological Society.

Rice, R. C., Sable, E. G., Denver, G. R., and Kehn, T. M., 1979, the Mississippian and Pennsylvanian (Carboniferous) systems in Kentucky: U.S. Geological Survey Prof. Paper 1110-F, p. F1-F32.

Rich, J. L., 1934-, Mechanics of low-angle overthrust faulting as illustrated by Cumberland thrust block, Virginia, Kentucky and Tennessee: Am. Assoc. Petroleum Geologists Bull., v. 18, no. 12, p. 1594--1596.

Rodgers; J., 194-9, Evolution of thought on structure of middle and southern Appalachians: Am. Assoc. Petroleum Geologists Bull., v. 38, no. 10, p. 164-3-1654-.

1963, Mechanics of Appalachian foreland folding in Pennsylvania and West Virginia: Am. Assoc. Petroleum Geologists BuU., 47, no. 8, p. 1527-1536.

1964-, Basement and No-Basement Hypotheses in Jura and the Appalachian Valley and Ridge: In Tectonics of the Southern Appalachians: Virginia Polytechnic Institute, Department of Geologica1.Sciences, Memoir 1, p. 71-80.

1972, Evaluation of though on structure of middle and southern Appalachians: Second paper, in Appalachians structures, origin, evaluation and possible potential for new exploration frontiers, a seminar, West Virginia University and West Virginia Geological and 99

Economic Survey, p. 1-15.

Ryan, \V. M., 1974, Structure and Hydrocarbon Production Associated with Pine Mountain Thrust System in Western Virginia (abs): Am. Assoc. Petroleum Geologists Bull., v. 58, no. 9, p. 1894.

Sabins, F. F., 1978, R-emote -sensing principles and interpretation, San Francisco: W. H. Freeman and Company. Sabins, F. F. Blom, R., and Elachi, c., 1980, Expression of San Andreas Fault on Seasat Radar Image, A.A.P.G. Bull., v. 64, p. 619-62.8.

U.S. Geological Survey EROS Data Center, (1982), Landsat 4: Landsat I Data Users Notes, no. 23, p. 1-12.

Wentworth, C. K., 1921, Russell Fork fault: in the geology and coal resources of Dickenson County, Virginia: Virginia Geol. Survey Bull. 21, p. 53-67.

Wilson, J. T., 1966, Did the Atlantic close and then re-open?: Nature, v. 211, p. 676-681. /

APPENDIX

U.S. GEOLOGICAL SURVEY QUADRANGEL MAPS

USED FOR THIS STUDY

100 101

Alvord, D. C., 1971~ Geologic map .of the Br

1971, Geologic map of the Hellier quadrangle, Kentucky- Virginia, and part of the Clintwood quadrangle, Pike County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-950.

1971a, Geologic map of the Naugatkuck and Delbarton quadrangles, eastern Kentucky.: U.S. Geological Survey Geol. Quad. Map

Alvord, D. C., and Trent, V. A., Geologic map of the Williamson quadrangle: U.S. Geological Survey Geol. Quad. Map GQ-187.

Alvord, D. C., and Holbrook, C. H., 1965, Geologic map of the Pikeville quadrangle, Pike and Floyd Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-480.

Alvord, D.C., and Miller, R. L., 1972, Geological map of the Elkhorn City quadrangle, Kentucky - Virginia and part of the Harmon quadrangle, Pike County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ- 951.

Barr, J. L., and Arndt, H. H., 1968, Geologic map of the Dorton quadrangle, Pike County, Kentucky: U.S. Geol. Survey Geol. Quad. Map GQ-713.

Connor, C. W., and Flores, R. M., 1978, Geologic map of the Louisa quadrangle, Kentucky- West Virginia: U.S. Geological Survey Geol. Quad. Map GQ-1462.

Danilchik, W., 1976_, Geologic map of the Hindman quadrangle, Knott County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ- 1372.

--- 197?, Geologic map .of the Handshoe quadrangle, eastern Kentucky: · U.S.- Geol. Survey Geol. Quad. Map GQ-1372.

1977a, Geologic map of the Tiptop quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1410.

Danilchik, W.. , and Waldrop, H. A., 1978, -Geologic map of the Vest quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1441.

Hayes, P. T., 1977, Geologic map of the Sitka quadrangle, Johnson and Lawrence Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1398.

Hinrichs, E. N., 1978, Geologic map of the Noble quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1476.

Hinrichs, E. N., and Rice, c. L., 1976, Geologic map of the Kite ------

102

quadrangle, so~.:~theast-em Kentucky! U.S. Geological Sl.:lrvey Geol. Quad. Map GQ-1397.

Hinrichs, E. N., and Ping, R. G., 1978, Geological map of the Wayland quadrangle, Knott and Floyd Counties, Kentucky: U.S. Geolgocial Survey Geol. Quad.Map GQ-1LJ.51.

Huddle, J. W., and Englund, K. J., 1962, Geologic map of the Kermit quadrangle in Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-178.

----- 1962a, Geologic map of the Varney quadrangle in Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-180.

Jenkins, E. C., 1966, Geologic map of the Milo quadrangle, Kentucky - West Virginia and part of the Webb quadrangle in Kentucky: U.S. Geol. Survey Geol. Quad. Map GQ-659.

-- 1967, Geologic map of the Millard quadrangle, Pike County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-5LJ.3.

Lee, K. Y ., Danilchik, W., and Rice, C. L., 1977, Geologic map of tile Guage quadrangle, Breathitt and Magoffin Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-llJ.16.

McKay, E. J., and Alvor

Outerbridge, W. F., 1963, Geologic map of the Inez quadrangle, Kentucky: U.S. GeoJgoical Survey Geol. Quad. Map GQ-,22.6.

1964, Geologic map of the Offutt quadrangle, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-3LJ.8. ------1966, Geologic map of the Paintsville quadrangle, Johnson and Floyd Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-lJ.95.

1967, Geologic map of the Oil Springs quadrangle, eastern Kentucky~ U.S. Geological Survey Geol. Quad. Map GQ-586.

----- 1968, Geologic map of the David quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-720.

-- 1968a, Geologic map of parts of the Majestic Hurley and Wharncliffe quadrangles, Pike County, Kentucky: U.S. Geological Survey Geol. Quad. map GQ-7LJ.8.

--- 1975, Geological map of the Wheelwright quadrangle, southeastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1251.

Outerbridge, W. F., and Van Vloten, R., 1968, Geologic map of the 103

Jamboree quadrangle, Pike County, Kentucky: U.S.. Geological Survey Geol. Quad. Map GQ-227.

Rice, C. L., 1963, Geologic map of Thomas quadrangle, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-347.

----- 1964, Geologic map of the Lancer quadrangle, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-347.

-- 1965, Geologic map of the Harold quadrangle, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-441.

---- 1966, Geologic map of the Martin quadrangle, Floyd County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-563.

-- 1967, Geologic map of the Prestonsburg quadrangle, Floyd and Johnson Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-641.

1968, Geologic map of the Redbush quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-708.

---- 1968a, Geologic map of the McDoweU quadrangle, Floyd and Pike Counties, Kentucky: U.S. Geological Survey Geol. Map GQ-732.

---- 1969, Geologic map of the Ivyton quadrangle, eastern Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-801.

----- 1973, Geologic map of the Jenkins West quadrangle, Kentucky - Virginia: U.S. Geologia! Survey Geol. Quad. Map GQ-1126.

--- 1976, Geologic map of the Mayking quadrangle, Letcher and Knott Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ- 1309.

'Rice, c. L., Ping, R., and Barr, J. L., 1977, Geologic map of the Belfry quadrangle, Pike County, Kentucky, U.S. Geological Survey Geol. Quad. Map GQ-1369.

Sanchez, J. D., Alvord, D. C., and Hayes, P. T., 1978, Geologic map of the Richardson quadrangle, Lawrence and Johnson Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1460.

Seiders, V. M., 1964, Geologic map of the Hazard North quadrangle, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-344.

1965, Geologic map of the Carrie quadrangle, Kentucky.: U.S. Geological Survey Geol. Quad Map GQ-422.

Spengler, R. \V., 1977, Geologic map of the Salyersville South quadrangle, Magoffin and Breathitt Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-276. lOlt-

Trent, V. A., 1965, Geologic map of the Matewan quadrangle in Kentucky: U.S. Geological Survey Geol. Quad. map GQ-373.

Ward, D. W., 1978, Geologic map of the Adams quadrangle, Lawrence County, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ­ llt-89.

Wolcott, D. E., 197lt-, Geologic map of the Jenkins East quadrangle, Pike and Letcher Counties, Kentucky: U.S. Geological Survey Geol. Quad. Map GQ-1210.

Wolcott, D. E., and Jenkins, E. C., 1966, Geologic map of the Meta quadrangle, Pike County, Kentucky: U.S. Geological St.kvey Geol. Quad. Map GQ-lt-97. ( I