10.06.2013 Views

permo-carboniferous carbonate platforms and reefs - Lock Haven ...

permo-carboniferous carbonate platforms and reefs - Lock Haven ...

permo-carboniferous carbonate platforms and reefs - Lock Haven ...

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

PERMO-CARBONIFEROUS CARBONATE<br />

PLATFORMS AND REEFS<br />

Volc<br />

W T-16 T-463 T-5246 T-5056 T-10 E<br />

500 m<br />

V.E. = 10<br />

5000 m<br />

Volc<br />

Unit 1<br />

Unit 2<br />

Unit 3<br />

Top Lvis<br />

Tul<br />

Bob<br />

Rad<br />

TSS<br />

HSS<br />

TSS HSS TSS HSS TSS<br />

Bash_SSB<br />

Serp_SSB<br />

Lvis13_csb<br />

Lvis2_MFS<br />

Lvis_SSB<br />

Lvis1_MFS<br />

Edited by:<br />

WAYNE M. AHR<br />

Department of Geology <strong>and</strong> Geophysics, Texas A&M University, College Station, Texas 77843-3115, U.S.A.<br />

PAUL M. (MITCH) HARRIS<br />

ChevronTexaco E&P Technology Company, 6001 Bollinger Canyon Road,<br />

San Ramon, California 94583-0746, U.S.A.<br />

WILLIAM A. MORGAN<br />

ConocoPhillips, Inc., P.O. Box 2197, Houston, Texas 77252-2197, U.S.A.<br />

AND<br />

IAN D. SOMERVILLE<br />

Department of Geology, University College - Dublin, Belfield, Dublin 4, Irel<strong>and</strong><br />

Top Evis<br />

Evis_SSB<br />

Tour_MFS<br />

Fame_SSB<br />

Bash = Bashkirian<br />

Serp = Serpukhovian<br />

Tul = Tulsky<br />

Bob = Bobrikovsky<br />

Tour = Tournaisian<br />

Fame = Famennian (Devonian) Devonian<br />

Lvis = Late Visean Rad = Radaevsky SSB = Supersequence Boundary<br />

“Volc” = Thick Ash Interval Evis - Early Visean MFS = Maximum Flooding Surface<br />

TSS = Transgressive Sequence Set HSS = Highst<strong>and</strong> Sequence Set<br />

SEPM Special Publication 78<br />

AAPG Memoir 83<br />

Platform (Shallow & Deeper)<br />

Upper-Middle Slope &<br />

Deeper Platform<br />

Allochthonous Debris<br />

Lower Slope -<br />

Toe of Slope<br />

“Volc” Interval<br />

Basin<br />

Published jointly by<br />

SEPM (Society for Sedimentary Geology) <strong>and</strong><br />

The American Association of Petroleum Geologists (AAPG)<br />

Tulsa, Oklahoma, U.S.A.<br />

Printed in the United States of America<br />

December, 2003<br />

Volcanic


SEPM, AAPG, <strong>and</strong> the authors are grateful to the following<br />

for their generous contribution to the cost of publishing<br />

Permo-Carboniferous Carbonate Platforms <strong>and</strong> Reefs<br />

ExxonMobil Development Company<br />

Kansas Geological Survey<br />

Saudi Aramco<br />

University of Kansas, Department of Geology, Haas Fund<br />

Contributions were applied to the cost of production, which reduced the<br />

purchase price, making the volume available to a wide audience<br />

SEPM <strong>and</strong> AAPG grant permission for a single photocopy of an item from this publication for personal use.<br />

Authorization for additional copies of items from this publication for personal or internal use is granted by SEPM<br />

<strong>and</strong> AAPG provided that the base fee of $3.50 per copy <strong>and</strong> $0.50 per page is paid directly to the Copyright<br />

Clearance Center, 222 Rosewood Drive, Danvers, MA 01923 (phone 978-750-8400). Fees are subject to change. Any<br />

form of electronic or digital scanning or other digital transformation of portions of this publication into computerreadable<br />

<strong>and</strong>/or transmittable for personal or corporate use requires special permission from, <strong>and</strong> is subject to fee<br />

charges by the SEPM <strong>and</strong> AAPG.<br />

SEPM Editor of Special Publications: Laura J. Crossey AAPG Association Editor: John C. Lorenz<br />

SEPM Executive Director: Howard E. Harper, Jr. AAPG Geoscience Director: J.B. “Jack” Thomas<br />

SEPM Publications Coordinator: Kris A. Farnsworth<br />

This book <strong>and</strong> other SEPM <strong>and</strong> AAPG publications are available from:<br />

SEPM The AAPG Bookstore<br />

6128 E. 38th. Street, Suite 308 P.O. Box 979<br />

Tulsa, OK 74135-5814, U.S.A. Tulsa, OK 74101-0979, U.S.A.<br />

Telephone: 1-918-610-3361 or 1-800-865-9765 (U.S.A.) Telephone: 1-918-584-2555 or 1-800-364-AAPG (U.S.A.)<br />

Fax: 1-918-621-1685 Fax: 1-918-560-2652 or 1-800-898-2274 (U.S.A.)<br />

www.sepm.org www.aapg.org<br />

The American Association of Petroleum Geologists (AAPG) does not endorse<br />

or recommend any product or services that may be cited, used, or discussed in<br />

AAPG publications or in presentations at events associated with the AAPG.<br />

ISBN 1-56576-087-5<br />

Copyright © 2003 by<br />

SEPM (Society for Sedimentary Geology) <strong>and</strong> The American Association of Petroleum Geologists (AAPG)<br />

All Rights Reserved<br />

Printed in the United States of America<br />

Published December, 2003


MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

SEQUENCE RESPONSE OF A DISTAL-TO-PROXIMAL FORELAND RAMP TO<br />

GLACIO-EUSTASY AND TECTONICS: MISSISSIPPIAN, APPALACHIAN BASIN,<br />

WEST VIRGINIA–VIRGINIA, U.S.A.<br />

AUS AL-TAWIL<br />

Saudi Aramco, Dhahran, Saudi Arabia<br />

AND<br />

THOMAS C. WYNN, AND J. FRED READ<br />

Department of Geological Sciences, Virginia Tech, Blacksburg Virginia 24061, U.S.A.<br />

ABSTRACT: This paper evaluates the limits of using well-cuttings data <strong>and</strong> wireline logs in conjunction with limited core <strong>and</strong> outcrop data<br />

to generate a regional, high-resolution sequence stratigraphy for upper Mississippian (Chesterian) Greenbrier <strong>carbonate</strong>s, West Virginia,<br />

U.S.A. These data are then used to document the stratigraphic response of the distal to proximal forel<strong>and</strong> basin to tectonics <strong>and</strong><br />

Carboniferous glacio-eustasy during the transition into ice-house times. The major mappable sequences are fourth-order sequences, a few<br />

meters to over a hundred meters thick. They consist of updip red beds <strong>and</strong> eolianites, lagoonal muddy <strong>carbonate</strong>s, ooid grainstone <strong>and</strong><br />

skeletal grainstone–packstone shoal complexes, open-ramp skeletal wackestone, <strong>and</strong> slope–basinal laminated argillaceous <strong>carbonate</strong>s. The<br />

sequences are bounded downdip by lowst<strong>and</strong> s<strong>and</strong>stones <strong>and</strong> calcareous siltstones, <strong>and</strong> locally on the ramp by basal transgressive shales;<br />

only a few sequence boundaries are calichified, compared with updip sections in Kentucky, where caliche <strong>and</strong> breccias are common.<br />

Transgressive systems tracts range from thin units to others that constitute the lower half of the sequence. The highst<strong>and</strong> systems tracts<br />

contain significant grainstone units. Maximum flooding surfaces on the ramp slope occur at the base of slope or basinal facies that rest on<br />

lowst<strong>and</strong> to transgressive complexes, whereas on the ramp they occur beneath widespread grainstones that overlie nearshore shale or lime<br />

mudstone. In the Greenbrier succession, fourth-order sequences are arranged into weak third-order composite sequences bounded updip<br />

by red beds, <strong>and</strong> by lowst<strong>and</strong> s<strong>and</strong>s <strong>and</strong> oolite along the ramp slope. The composite sequences contain three to four fourth-order sequences.<br />

Correlation of the forel<strong>and</strong> basin units with third-order global sea-level curves, <strong>and</strong> with high-frequency sequences within the<br />

intracratonic Illinois Basin, shows that in spite of differential subsidence rates ranging from 2 to 30 cm/ky across the forel<strong>and</strong>, global third<strong>and</strong><br />

fourth-order sea-level changes whose amplitude increased with time were a strong influence on sequence development. Thrust-loadinduced<br />

differential subsidence of fault blocks of the forel<strong>and</strong> basement controlled the rapid basinward thickening of the depositional<br />

wedge, <strong>and</strong> modified the eustatic effects on the accumulating succession. Moderate-amplitude eustatic sea-level change <strong>and</strong> semiarid<br />

climate were the dominant causes of the widespread reservoirs of ooid grainstone <strong>and</strong> lowst<strong>and</strong> s<strong>and</strong>s. The overall stratigraphy suggests<br />

upward increase in amplitudes of sea-level change, <strong>and</strong> cooling, which likely records the initiation of Gondwana ice-sheet growth.<br />

INTRODUCTION<br />

This paper mainly uses well data (especially well cuttings <strong>and</strong><br />

wireline logs), along with limited outcrop data <strong>and</strong> core, to<br />

determine if they can be used to generate a high-resolution threedimensional,<br />

sequence stratigraphic framework for the Mississippian<br />

(Chesterian) <strong>carbonate</strong>s, from 0 up to 200 m thick, in the<br />

eastern, Appalachian forel<strong>and</strong> basin (Colton, 1970; de Witt <strong>and</strong><br />

McGrew, 1979) (Figs. 1–4). The study differs from Al-Tawil <strong>and</strong><br />

Read (this volume), which was largely an outcrop-based study<br />

along two outcrop belts in the distal forel<strong>and</strong>, Kentucky. Its<br />

forel<strong>and</strong>-basin setting <strong>and</strong> presence of the ramp margin in the<br />

study area distinguish it from the studies of Smith <strong>and</strong> Read<br />

(1999, 2000, 2001) in the Illinois Basin, an intracratonic basin to the<br />

west in which the ramp margin is not present. The units in the<br />

study area show extremely complex facies heterogeneity, with<br />

rapid vertical <strong>and</strong> lateral facies changes that have hampered<br />

development of regional exploration models, <strong>and</strong> recognition of<br />

shoreline <strong>and</strong> ramp-margin trends.<br />

The paper provides the first detailed, high-resolution, regional<br />

sequence stratigraphic study of the mixed <strong>carbonate</strong> <strong>and</strong><br />

siliciclastic facies extending from the slowly subsiding distal<br />

forel<strong>and</strong> to the more rapidly subsiding proximal forel<strong>and</strong> of the<br />

Appalachian Basin. Previously, the succession had been subdivided<br />

into several lithologically heterogeneous formations (Fig.<br />

4), with much of the correlation being based on a few semiregional<br />

siliciclastic marker beds (Reger, 1926; Wells, 1950;<br />

McFarlan <strong>and</strong> Walker, 1956; Rice et al., 1979). We subdivided the<br />

Chesterian succession into regionally traceable fourth-order high-<br />

Permo-Carboniferous Carbonate Platforms <strong>and</strong> Reefs<br />

SEPM Special Publication No. 78 <strong>and</strong> AAPG Memoir 83, Copyright © 2003<br />

SEPM (Society for Sedimentary Geology), American Association of Petroleum Geologists (AAPG), ISBN 1-56576-087-5, p. 11–34.<br />

11<br />

frequency depositional sequences, by tracing major l<strong>and</strong>ward<br />

<strong>and</strong> basinward shifts in facies, <strong>and</strong> mapping outcropping sequence-bounding<br />

disconformities, correlative conformities, <strong>and</strong><br />

sequence-bounding siliciclastic units. This largely subsurface<br />

study illustrates how high-resolution sequence stratigraphic<br />

analysis can increase our underst<strong>and</strong>ing of facies stacking on the<br />

ramp (Ahr, 1973) <strong>and</strong> its response to forel<strong>and</strong> basin tectonics,<br />

glacio-eustatic sea-level changes related to onset of Gondwana<br />

glaciation (Crowell, 1978), <strong>and</strong> long-term climate change<br />

(Ettensohn et al., 1988).<br />

The study documents how, even in this tectonically active<br />

forel<strong>and</strong>-basin setting, the major role of tectonics has been one of<br />

generating differential accommodation across <strong>and</strong> along the<br />

forel<strong>and</strong> basin, <strong>and</strong> in localizing the position of structural highs,<br />

the ramp’s hinge line, <strong>and</strong> ramp-margin facies. In spite of the<br />

active tectonic regime, fourth-order <strong>and</strong>, to a lesser extent, thirdorder<br />

eustasy has been the overriding control on the generation<br />

of the regionally correlative sequences.<br />

GEOLOGIC SETTING<br />

Regional Structural Setting<br />

The Appalachian Basin is an elongate forel<strong>and</strong> basin bordered<br />

on the east by overthrusted Precambrian <strong>and</strong> early Paleozoic rocks<br />

<strong>and</strong> on the west by the Cincinnati Arch (Colton, 1970; de Witt <strong>and</strong><br />

McGrew, 1979) (Figs. 1, 2). In the west, the Paleozoic rocks are<br />

undeformed to gently folded, whereas the eastern part of the<br />

basin is in the folded <strong>and</strong> thrusted Valley-<strong>and</strong>-Ridge Province.


12<br />

The Upper Mississippian <strong>carbonate</strong> strata in the Appalachian<br />

Basin (Arkle et al., 1979) are exposed in narrow, northeast–<br />

southwest trending outcrop belts, one along the distal forel<strong>and</strong> in<br />

Kentucky adjacent to the Cincinnati Arch (described by Al-Tawil<br />

<strong>and</strong> Read, this volume) <strong>and</strong> the others (described here) in the<br />

more proximal forel<strong>and</strong>, exposed along the eastern Appalachian<br />

Plateau <strong>and</strong> into the fold–thrust belt (Figs. 1, 2). Between these<br />

outcrop belts, the Mississippian rocks of the distal to proximal<br />

forel<strong>and</strong> basin lie in the subsurface beneath Pennsylvanian sediments<br />

<strong>and</strong> are penetrated by numerous wells, twenty of which<br />

were used along with a core <strong>and</strong> four outcrop sections for this<br />

study.<br />

The Appalachian forel<strong>and</strong> basin was tectonically active during<br />

the deposition of these Upper Mississippian <strong>carbonate</strong>s, <strong>and</strong><br />

the Mississippian rocks thicken into the foredeep in southwestern<br />

Virginia <strong>and</strong> Tennessee (Fig. 1). The depocenter of the Late<br />

Mississippian <strong>carbonate</strong> ramp is located in the overthrust belt in<br />

the Greendale Syncline, southwestern Virginia, where Mississippian<br />

units thicken to over 2000 meters of basinal, slope, <strong>and</strong><br />

ramp-margin facies (Cooper, 1948; Bartlett <strong>and</strong> Webb, 1971).<br />

Several synsedimentary positive structures occur in West Virginia<br />

<strong>and</strong> eastern Kentucky (Yeilding <strong>and</strong> Dennison, 1986; Dever<br />

et al., 1990; Sable <strong>and</strong> Dever, 1990; Dever, 1995) (Fig. 1). The<br />

Cincinnati Arch, which borders the western edge of the basin,<br />

subsided very little during Chesterian time, while adjacent<br />

areas to the west (Illinois Basin) <strong>and</strong> the east (Appalachian<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

FIG. 1.—Geologic location map of study area, showing distribution of exposed Mississippian Greenbrier <strong>carbonate</strong> rocks (gray<br />

shading) in the Appalachian Basin, eastern U.S.A. Isopach contours (in meters) show thickness of Greenbrier Group, which<br />

thickens into the Appalachian foredeep to the southeast (modified from Flowers, 1956; Pryor <strong>and</strong> Sable, 1974; MacQuown <strong>and</strong><br />

Pear, 1980; Yeilding <strong>and</strong> Dennison, 1986; Dever, 1986, 1995; Sable <strong>and</strong> Dever, 1990, Dever et al., 1990). Major structures within the<br />

study area are the ancestral Rome Trough border faults, <strong>and</strong> the Jessamine Dome <strong>and</strong> Cincinnati <strong>and</strong> Waverly arches, which<br />

bound the area to the northwest, <strong>and</strong> the Burning–Mann anticline <strong>and</strong> Virginia Promontory, which cut across the trend of the<br />

basin. Orogenic highl<strong>and</strong>s bounded the basin to the southeast.<br />

Basin) subsided more rapidly (Pryor <strong>and</strong> Sable, 1974; deWitt<br />

<strong>and</strong> McGrew, 1979; Whitehead, 1984; Sable <strong>and</strong> Dever, 1990;<br />

Dever, 1995). The West Virginia Dome was a positive structure<br />

along the 38th parallel lineament in northeastern West Virginia,<br />

which was active during Early Chesterian time (Yeilding <strong>and</strong><br />

Dennison, 1986).<br />

The basement in the study area is segmented by major faults<br />

into a series of blocks, some of which moved independently<br />

during Mississippian deposition (Shumaker <strong>and</strong> Wilson, 1996).<br />

These faults include northeast–southwest trending bounding<br />

faults of the Rome Trough, as well as several zones parallel to <strong>and</strong><br />

to the east of these. The northeast–southwest trending basinal<br />

hinge occurs just to the west of the overthrust belt, <strong>and</strong> a smaller<br />

east–west trending hinge occurs in the northeast along the 400foot<br />

(120 m) Greenbrier isopach (Fig. 1). A northwest–southeast<br />

trending arch lies at right angles to the regional structure, in the<br />

vicinity of the “Virginia Promontory”, a long-lived Paleozoic<br />

high (Thomas, 1977).<br />

Stratigraphy <strong>and</strong> Biostratigraphy<br />

The Mississippian succession makes up a supersequence set<br />

(roughly 35 My duration) composed of at least three<br />

supersequences. The lower supersequence set boundary is the<br />

Devonian–Mississippian erosional unconformity (Bjerstadt <strong>and</strong><br />

Kammer, 1988) (Figs. 3, 4). The lower supersequence consists of


WAVERLY<br />

ARCH<br />

ROME ROME TROUGH TROUGH<br />

APPALACHIAN<br />

BASIN<br />

OHIO<br />

the up<strong>permo</strong>st Devonian–lowermost Mississippian Cloyd Conglomerate–Berea<br />

S<strong>and</strong>stone <strong>and</strong> the overlying Subury Shale–<br />

Price/Borden/Maccrady deltaic complex (Kinderhookian to<br />

mid-Osagean), which contains several clinoformed depositional<br />

sequences (Branson, 1912; Butts, 1940, 1941; Bjerstadt <strong>and</strong><br />

Kammer, 1988; Bartlett, 1974). This is overlain southwest of the<br />

study area by the <strong>carbonate</strong>-prone Fort Payne–Salem<br />

supersequence (upper Osagean–lower Meramecian) (Fig. 4;<br />

Bjerstadt <strong>and</strong> Kammer, 1988; Sable <strong>and</strong> Dever, 1990; Khetani<br />

<strong>and</strong> Read, 2002).<br />

The upper Meramecian–Chesterian supersequence (Fig. 4)<br />

described in this paper overlies the lower two supersequences.<br />

The upper Meramecian Little Valley–Hillsdale sequence may<br />

have formed during lowered sea levels relative to the overlying<br />

units, given that it is the least extensive unit on the platform, <strong>and</strong><br />

it contains relatively shallow oncolitic limestone facies even in the<br />

most basinward sections. The Chesterian sequences C1 to 11<br />

described in this paper make up the <strong>carbonate</strong>-prone, largely<br />

transgressive part of the supersequence. The shale-prone sequence<br />

C11 may be the initial sequence of the supersequence<br />

highst<strong>and</strong>, the bulk of the supersequence highst<strong>and</strong> being made<br />

up of the progradational Mauch Chunk red siliciclastics. The<br />

early Pennsylvanian unconformity marks the top boundary of<br />

the supersequence set.<br />

Regional Mississippian stratigraphy <strong>and</strong> biostratigraphy of<br />

the Mississippian in Virginia <strong>and</strong> West Virginia (Figs. 3, 4) are<br />

given in Butts (1922, 1940, 1941), Reger (1926), Wells (1950),<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

40° N 83° W 40° N 80° W<br />

11<br />

BURNING-MANN<br />

ANTICLINE<br />

FOLD AND THRUST BELT<br />

WEST<br />

VIRGINIA<br />

12<br />

14<br />

13<br />

15<br />

17<br />

16<br />

18 20<br />

5<br />

19<br />

21<br />

22<br />

24<br />

9<br />

23<br />

25 26<br />

10<br />

VIRGINIA<br />

PROMONTORY<br />

WEST VIRGINIA<br />

DOME<br />

FIG. 2.—Location map showing outcrop sections <strong>and</strong> well sections used to construct the cross sections in the study. Location<br />

information on sections is in the Appendix. Outcrops of Mississippian Greenbrier Limestone is shaded. The northeast–southwest<br />

cross section, based on sections 1 to 11 (Fig. 6), includes data from the eastern outcrop belt, supplemented by nearby shallow wells<br />

<strong>and</strong> a core; the sections thicken southwestward down the basin axis. The northwest–southeast cross section extends across the<br />

forel<strong>and</strong> normal to the trend of the basin <strong>and</strong> is based on wells 12 to 26 (Fig. 7); it uses only subsurface well cuttings <strong>and</strong> wireline<br />

logs. The succession thickens markedly across the northeast–southwest trending hinge (just southeast of the 100 m isopach of<br />

Figure 1) that separates the distal forel<strong>and</strong> from the proximal forel<strong>and</strong>. Note that to the northeast this hinge curves to the east.<br />

Two minor east–west trending hinges occur to the southwest of the main hinge. Outcrops representative of the distal forel<strong>and</strong><br />

of Kentucky (along strike to the southwest) are described in Al-Tawil <strong>and</strong> Read (this volume).<br />

4<br />

7<br />

8<br />

DISTAL FORELAND<br />

6<br />

2<br />

3<br />

1<br />

37° N 80° W<br />

VIRGINIA<br />

PROXIMAL FORELAND<br />

0<br />

100 km<br />

100 mi<br />

MISSISSIPPIAN<br />

OROGENIC<br />

HIGHLANDS<br />

13<br />

Flowers (1956), de Witt <strong>and</strong> McGrew (1979), Rice et al. (1979), <strong>and</strong><br />

Maples <strong>and</strong> Waters (1987). The Greenbrier Group (Hillsdale to<br />

Alderson interval <strong>and</strong> the Lillydale Shale <strong>and</strong> Glenray Limestone<br />

of the Lower Bluefield Formation, Mauch Chunk Group) make<br />

up the study interval (0 to 1000 m thick). It unconformably<br />

overlies the Price–Maccrady deltaic–marine shelf deposits (Figs.<br />

3, 4). The basal unit studied is the upper Meramecian Little<br />

Valley–Hillsdale Limestone, which is a very cherty, pelletal–<br />

peloidal <strong>and</strong> skeletal packstone–wackestone with the distinctive<br />

corals Syringopora <strong>and</strong> Acrocyathus proliferus (Lithostrotionella)<br />

(Wells, 1950; DiRenzo, 1986). This is overlain by the Chesterian<br />

Denmar Formation, which in outcrop contains numerous exposure<br />

surfaces, consists of ooid grainstone, lime wackestone–<br />

mudstone, <strong>and</strong> minor quartz–peloidal grainstone, <strong>and</strong> has the<br />

crinoid Platycrinites penicillus <strong>and</strong> the earliest occurrence of the<br />

zonal conodonts G. bilineatus <strong>and</strong> C. charactus (Butts, 1922, 1940,<br />

1941; Wells, 1950; McFarlan <strong>and</strong> Walker, 1956; Dever et al., 1990).<br />

The Denmar Formation is overlain by the Taggard red beds, a<br />

regional marker consisting of red shale <strong>and</strong> siltstone <strong>and</strong> minor<br />

<strong>carbonate</strong>s.<br />

The overlying ooid- <strong>and</strong> skeletal limestone–rich units <strong>and</strong><br />

subordinate <strong>carbonate</strong> mudstone constitute the Gasper Formation,<br />

which contains the first appearance of the crinoid index<br />

fossil Talarocrinus (Butts, 1940, 1941; Wells, 1950). The interval is<br />

subdivided into the Pickaway, Union, Greenville Shale, <strong>and</strong><br />

Alderson Formations (Reger, 1926). This is overlain by Lillydale<br />

Shale <strong>and</strong> the Glenray Limestone (Reger, 1926), consisting of


14<br />

PERIOD<br />

MISSISSIPPIAN<br />

DEV.<br />

EPOCH<br />

KIND. OSAGEAN MERAM. CHESTERIAN<br />

AGE<br />

327<br />

354<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

CONODONT<br />

ZONES<br />

G. bilineatus -<br />

K. mehli<br />

G. bilineatus -<br />

C. altus<br />

G. bilineatus -<br />

C. charactus<br />

A. scalenus -<br />

Cavusgnathus<br />

Cavusgnathus -<br />

T. varians<br />

OTHER<br />

INDEX<br />

FOSSILS<br />

Talarocrinus<br />

spp.<br />

Platycrinites<br />

penicillus<br />

Acrocyathus spp.<br />

'Lithostrotion spp.'<br />

MAUCH CHUNK GP.<br />

G R E E N B R I A R G R O U P ("B I G L I M E")<br />

WEST VIRGINIA<br />

DROOP SS.<br />

REYNOLDS<br />

LIMESTONE<br />

GLENRAY<br />

LIMESTONE<br />

LILLYDALE<br />

SHALE<br />

ALDERSON<br />

LIMESTONE<br />

UNION<br />

LIMESTONE<br />

PICKAWAY<br />

LIMESTONE<br />

UPPER / LOWER<br />

TAGGARD<br />

DENMAR<br />

LIMESTONE<br />

VIRGINIA<br />

COVE<br />

CREEK<br />

LIMESTONE<br />

FIDO SS.<br />

HILLSDALE LST.<br />

LITTLE VALLEY LST.<br />

MACCRADY<br />

FORMATION<br />

PRICE<br />

FORMATION<br />

PENNINGTON<br />

FM.<br />

GASPER<br />

LIMESTONE<br />

STE.<br />

GENEVIEVE<br />

LIMESTONE<br />

LATE DEVONIAN AND<br />

EARLY MISSISSIPPIAN<br />

BLACK SHALE<br />

SEQ.<br />

THIS<br />

PAPER<br />

FIG. 3.—Chronostratigraphic chart, Mississippian interval, showing the major units in the study area of West Virginia <strong>and</strong> Virginia;<br />

studied interval is from the Little Valley–Hillsdale to the Glenray Formation, <strong>and</strong> includes the drillers’ “Big Lime” (Greenbrier<br />

Group) <strong>and</strong> the “Little Lime”. Chesterian sequences described in this paper are numbered C1 to C11; underlying Meramecian<br />

sequences are not numbered. The study interval is equivalent to the St. Louis to Glen Dean interval of Kentucky (Appalachian<br />

Basin) <strong>and</strong> the Illinois Basin. Modified from de Witt <strong>and</strong> McGrew (1979), Roberts et al (1995), Collinson et al (1971), Rexroad <strong>and</strong><br />

Clarke (1960), Huggins (1983), Stamm (1997), Barlett <strong>and</strong> Webb (1971), Reger (1926), <strong>and</strong> Wells (1950).<br />

ooid <strong>and</strong> skeletal limestone <strong>and</strong> containing the zonal conodonts<br />

G. bilineatus <strong>and</strong> Kladognathus (Rexroad <strong>and</strong> Clarke, 1960). The<br />

succession is capped by Mauch Chunk Group red s<strong>and</strong>stone,<br />

siltstone, <strong>and</strong> mudrock, whose top is the Mississippian–Pennsylvanian<br />

unconformity (Brezinski, 1989). Regionally, the units<br />

thicken to over 600 m to the southeast into the foredeep, where<br />

they are dominated by thick units of dark skeletal wackestone–<br />

mudstone <strong>and</strong> basin–slope laminated argillaceous <strong>carbonate</strong><br />

mudstones, <strong>and</strong> thin quartz s<strong>and</strong>stone <strong>and</strong> skeletal–ooid<br />

grainstone. In the subsurface, the Hillsdale to Alderson <strong>carbonate</strong>s<br />

are the drillers’ “Big Lime” <strong>and</strong> the Glenray Limestone is<br />

the “Little Lime.”<br />

BLUEFIELD FM.<br />

C11<br />

C10<br />

C9<br />

C8<br />

C7<br />

C6<br />

C5<br />

C4<br />

C3<br />

C2<br />

C1<br />

FACIES AND REGIONAL DISTRIBUTION<br />

The facies developed in the eastern Appalachian Basin<br />

(Leonard, 1968; Carney <strong>and</strong> Smosna, 1989) are shown schematically<br />

on an idealized ramp model (Fig. 5) <strong>and</strong> resemble facies<br />

described elsewhere by Ettensohn et al. (1984) <strong>and</strong> Smith <strong>and</strong><br />

Read (1999, 2001). The regional facies distributions are shown on<br />

the two regional cross sections (Figs. 6A, 7A). The first cross<br />

section extends southwestward down the axis of the forel<strong>and</strong><br />

basin, <strong>and</strong> is based on five detailed outcrop sections, a core, <strong>and</strong><br />

data from four wells. The second cross section is a dip section<br />

extending across the distal forel<strong>and</strong> onto the proximal forel<strong>and</strong>,


PERIOD<br />

MISSISSIPPIAN<br />

DEV.<br />

EPOCH<br />

KIND. OSAGEAN MERAM. CHESTERIAN<br />

AGE<br />

327<br />

354<br />

DISTAL PROXIMAL<br />

FORELAND<br />

MAUCH CHUNK<br />

SILICICLASTICS<br />

REYNOLDS LST.<br />

GLENRAY LST.<br />

LILLYDALE<br />

SHALE<br />

U. TAGGARD<br />

L. TAGGARD<br />

ALDERSON LST.<br />

UNION LST.<br />

PICKAWAY LST.<br />

R R R<br />

R R<br />

DENMAR LST.<br />

PRICE FM.<br />

SILICICLASTICS<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

COVE CREEK LST.<br />

HILLSDALE LST.<br />

L. VALLEY LST.<br />

MACCRADY<br />

FM.<br />

LATE DEVONIAN AND EARLY<br />

MISSISSIPPIAN BLACK SHALE<br />

FIDO SS.<br />

FIG. 4.—Schematic lithostratigraphy showing the major Mississippian<br />

stratigraphic units in the area. The study interval<br />

(Little Valley–Hillsdale to the Glenray/Cove Creek Limestone),<br />

overlies Early Mississippian siliciclastics of the Price–<br />

Borden formations, <strong>and</strong> is overlain by <strong>and</strong> interfingers with<br />

the upper Mississippian Mauch Chunk siliciclastics.<br />

PALEOSOLS<br />

QUARTZ-PELOID<br />

GRAINSTONE / PACKSTONE<br />

RED BEDS<br />

SEQUENCE BOUNDARY<br />

PELOIDAL<br />

GRAINSTONE / PACKSTONE<br />

LIME WACKESTONE / MUDSTONE<br />

~ 200 km<br />

OOID<br />

GRAINSTONE / PACKSTONE<br />

GRAY SHALE/SILTSTONE<br />

15<br />

<strong>and</strong> consists solely of data from 15 wells (cuttings with 10 to 15<br />

foot sample intervals <strong>and</strong> wire-line logs). Figures 6B <strong>and</strong> 7B show<br />

the individual sequences from each cross section rehung using<br />

the top of each sequence as a datum, <strong>and</strong> using inferred water<br />

depths of the topmost facies to hang the ramp-margin sections.<br />

Facies are summarized in Tables 1 <strong>and</strong> 2.<br />

Paleosols <strong>and</strong> Erosion Surfaces<br />

Carbonate paleosols occur on disconformities on the distal<br />

forel<strong>and</strong>, southwest of the study area in Kentucky (Dever, 1973;<br />

Ettensohn, 1975; Harrison <strong>and</strong> Steinen, 1978; Al-Tawil, 1998) <strong>and</strong><br />

West Virginia, but they are less common in the thicker West<br />

Virginia–Virginia units (Figs. 6, 7). The paleosols include laminar<br />

caliches <strong>and</strong> caliche breccia (Table 1). They formed under seasonally<br />

wet–dry semiarid conditions (Ettensohn et al., 1988). However,<br />

many contacts between sequences <strong>and</strong> parasequences in<br />

outcrop on the proximal forel<strong>and</strong> in West Virginia are sharp<br />

planar erosional to scalloped surfaces. They resemble microkarst<br />

or tidal solution basins.<br />

Red Beds (Marginal Marine)<br />

These are red, commonly structureless, unfossiliferous, fine<br />

siliciclastic mudrocks <strong>and</strong> laminated shale (Figs. 6, 7; Table 1).<br />

Updip in West Virginia, ripple lamination, wave lamination, <strong>and</strong><br />

mud cracks are common, but burrows are rare. Some red beds<br />

have rooted horizons, <strong>and</strong> some layers are burrowed. The red<br />

beds are common in updip areas, where they are associated with<br />

eolianites, <strong>and</strong> restricted lagoonal <strong>and</strong> shoal-water grainstones.<br />

These updip facies are marginal marine to intertidal–supratidal<br />

red beds, which formed as highst<strong>and</strong>, basinward-shifting facies.<br />

Gray Shale, Quartz S<strong>and</strong>stone, <strong>and</strong> Calcareous Siltstone<br />

(Tidally Influenced Open Siliciclastic Shelf)<br />

These facies include dark gray, poorly fissile, limy shales that<br />

are barren to variably fossiliferous, with small productid brachiopods,<br />

pelecypods, <strong>and</strong> bryozoa (Table 1). Some overlie pyritized<br />

bedding planes with abundant lacy bryozoans, especially at the<br />

tops of sequences C10 <strong>and</strong> C11 (Alderson <strong>and</strong> Glenray Formations).<br />

The shales form units a few centimeters thick that pinch<br />

out laterally, or are extensive sheets up to 30 m thick (e.g., in<br />

SKELETAL<br />

GRAINSTONE / PACKSTONE<br />

CALCAREOUS SILTSTONE<br />

SANDSTONE<br />

ARGILLACEOUS SKELETAL<br />

WACKESTONE / PACKSTONE<br />

LAMINATED SHALY<br />

LIME MUDSTONE /CALCAREOUS SILTSTONE<br />

FIG. 5.—Schematic facies profile for the Mississippian of the Appalachian Basin. Actual facies distributions are more complex, in that<br />

skeletal grainstone–packstone <strong>and</strong> ooid grainstone facies are developed not only on the ramp margin but also in local areas far<br />

back in the ramp interior.<br />

0 m<br />

20 m<br />

60 m


16<br />

150 MILES<br />

1 2 3 4 5 6 7 8 9 10 11<br />

CANAAN RANDOLPH<br />

GREENBRIER SUMMERS SUN OIL GREENBRIER<br />

MERCER<br />

VALLEY<br />

28 MINGO<br />

21<br />

7<br />

CORE<br />

15 UNION<br />

25 INGLESIDE HOLSTON<br />

FORMATION<br />

SEQ.<br />

GLENRAY LST.<br />

GREENVILLE<br />

SHALE<br />

BLUEFIELD<br />

FM.<br />

COVE CREEK<br />

LIMESTONE<br />

C11<br />

LILLYDALE SHALE<br />

DATUM<br />

FIDO SS.<br />

ALDERSON<br />

FM.<br />

C10<br />

C9<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

UNION<br />

FM.<br />

C8<br />

RED-BEDS<br />

PICKAWAY<br />

FM.<br />

GRAY SHALE / SILTSTONE<br />

C7<br />

CALCAREOUS SILTSTONE<br />

C6<br />

QUARTZ SANDSTONE<br />

TAGGARD<br />

FM.<br />

C5<br />

PRICE / MACCRADY FM.<br />

QUARTZ PELOID GRAINSTONE<br />

FINE-GRAINED PELLETAL PACKSTONE,<br />

LIME WACKESTONE / MUDSTONE<br />

OOID GRAINSTONE / PACKSTONE<br />

C4<br />

PELOID GRAINSTONE / PACKSTONE<br />

SKELETAL GRAINSTONE / PACKSTONE, MINOR WACKESTONE<br />

DENMAR<br />

FM.<br />

C3<br />

ARGILLACEOUS SKELETAL PACKSTONE / WACKESTONE<br />

200'<br />

LAMINATED SHALY LIME MUDSTONE/<br />

CALCAREOUS SILTSTONE<br />

SEQUENCE BOUNDARY<br />

C2<br />

PALEOSOL<br />

C1<br />

HILLSDALE<br />

FM.<br />

FIG. 6.—A) High-resolution sequence stratigraphic cross section along the eastern outcrop belt (parallel to the axis of the forel<strong>and</strong> basin), incorporating outcrop sections<br />

<strong>and</strong> nearby shallow wells, which were analyzed using cuttings <strong>and</strong> wireline logs. The cross section extends from West Virginia into Virginia, <strong>and</strong> was constructed<br />

by hanging sections from the regional Lillydale Shale marker; it shows the interpreted regional distribution of lithofacies, sequence boundaries, <strong>and</strong> their relation to<br />

the previous formational stratigraphy (modified from Al-Tawil, 1998). Legend is used for Figure 6B <strong>and</strong> 7B). Detailed information on well <strong>and</strong> outcrop locations is<br />

given in the Appendix.


MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

1 2 3 4 5 6 7 8 9 11<br />

1 2 3 4 5 6 7 8 9 10 11<br />

100'<br />

FEET<br />

0 MILES 10<br />

100'<br />

0 MILES 10<br />

HILLSDALE<br />

FIG. 6.—B) Facies distributions within individual sequences in Figure 6A (axial cross section), constructed by rehanging each<br />

sequence from its top sequence boundary, which was assumed to be horizontal on the ramp, <strong>and</strong> allowing the ramp margin to<br />

slope into the basin so that water depths of facies are compatible with estimated water depths in the text. Dotted line shows<br />

estimated positions of maximum flooding surfaces. Legend is same as Figure 6A.<br />

FEET<br />

C5<br />

C4<br />

C3<br />

C2<br />

C1<br />

C11<br />

C10<br />

C9<br />

C8<br />

C7<br />

C6<br />

17


18<br />

23 24 25 26<br />

FORMATION<br />

SEQ.<br />

22<br />

21<br />

20<br />

?<br />

C12<br />

19<br />

18<br />

BLUEFIELD<br />

FM.<br />

17<br />

?<br />

GREENVILLE<br />

SHALE<br />

16<br />

15<br />

DROOP SS.<br />

13 14<br />

12<br />

C11<br />

?<br />

LILLYDALE SHALE<br />

DATUM<br />

ALDERSON<br />

FM.<br />

C10<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

UNION<br />

FM.<br />

C9<br />

C8<br />

100<br />

PICKAWAY<br />

FM.<br />

FEET<br />

C7<br />

10<br />

20 = WYOMING 253<br />

21 = WYOMING 87<br />

22 = MCDOWELL 63<br />

23 = MCDOWELL 792<br />

24 = MERCER 16<br />

25 = MERCER 1<br />

26 = MERCER 27<br />

MILES<br />

0<br />

12 = WAYNE 277<br />

13 = WAYNE 1497<br />

14 = LINCOLN 155<br />

15 = LINCOLN 1375<br />

16 = MINGO 796<br />

17 = LOGAN 102<br />

18 = LOGAN 922<br />

19 = LOGAN 388<br />

12 13 14 15 16 17 18 19 20 21 22 23 24 25<br />

C6<br />

TAGGARD<br />

EQUIVALENT<br />

DATUM<br />

C5<br />

?<br />

?<br />

C4<br />

OOID GRAINSTONE / PACKSTONE<br />

RED BEDS<br />

DENMAR<br />

FM.<br />

C3<br />

PRICE / MACCRADY FM.<br />

PELOID GRAINSTONE / PACKSTONE<br />

GRAY SHALE / SILTSTONE<br />

SKELETAL GRAINSTONE / PACKSTONE,<br />

MINOR WACKESTONE<br />

CALCAREOUS SILTSTONE<br />

QUARTZ SANDSTONE<br />

C2<br />

ARGILLACEOUS SKELETAL<br />

PACKSTONE / WACKESTONE<br />

LAMINATED SHALY LIME MUDSTONE/<br />

CALCAREOUS SILTSTONE<br />

QUARTZ-PELOID GRAINSTONE<br />

C1<br />

FINE-GRAINED PELLETAL<br />

PACKSTONE,LIME<br />

WACKESTONE / MUDSTONE<br />

SEQUENCE BOUNDARY<br />

DOLOMITE<br />

HILLSDALE<br />

FM.<br />

PALEOSOL<br />

FIG. 7.—A) High-resolution sequence stratigraphic cross section from updip to downdip in West Virginia based on subsurface wells analyzed using cuttings <strong>and</strong> wireline<br />

logs (based on Thomas Wynn’s Ph.D. study, in prep.). Upper units were hung from the Lillydale Shale marker, as in Figure 6, but the lower sequences were hung from<br />

the base of the upper Taggard equivalent (quartz–peloid eolianites–calcareous siltstones at base of sequence C6). Cross section shows the interpreted regional<br />

distribution of lithofacies, sequence boundaries, <strong>and</strong> their relation to the previous formational stratigraphy. Note the similarity of the sequence stratigraphy <strong>and</strong> facies<br />

distributions of this well-based cross section to the dominantly outcrop- <strong>and</strong> well-based cross section of Figure 6. Detailed well location information is in the Appendix.


MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

12 13 14 15 16 17 18 19 20 21 22 23 24 25 26<br />

100 FT. 30 M<br />

16 KM.<br />

0 10 MILES<br />

FIG. 7.—B) Facies distributions within individual sequences in Figure 7A, constructed by rehanging each sequence from its top<br />

sequence boundary, which was assumed to be horizontal. Legend is same as Figure 7A. Dotted line shows estimated positions<br />

of maximum flooding surfaces.<br />

sequence C11) that grade up into argillaceous, skeletal limestones<br />

(Figs. 6, 7). S<strong>and</strong>stone <strong>and</strong>/or calcareous siltstone forms units up<br />

to 10 meters thick along the ramp margin <strong>and</strong> slope. Calcareous<br />

siltstones <strong>and</strong> s<strong>and</strong>stones are festoon cross-bedded to waveripple<br />

cross laminated or massive in outcrop <strong>and</strong> generally are<br />

unfossiliferous (Table 1).<br />

The shales are quiet-water deposits. Poorly fossiliferous thin<br />

shales formed in lagoons shoreward of s<strong>and</strong>stone–siltstone units<br />

at bases of sequences (cf. Ettensohn et al., 1984). However, thick<br />

fossiliferous shales such as those in sequence C11 may be deeperramp,<br />

open marine units (cf. Ettensohn et al., 1984). High<br />

siliciclastic influx may have been due to more humid climate,<br />

19<br />

perhaps coupled with short-term sea-level lowering. This increased<br />

turbidity suppressed <strong>carbonate</strong> productivity on the ramp,<br />

generating widespread shale deposits. The calcareous siltstone<br />

contains sedimentary structures that suggest a high-energy wave<strong>and</strong><br />

current-influenced coastal setting during shallowing phases<br />

on the outer ramp.<br />

Quartz–Peloid Grainstone<br />

(Coastal Eolianites, Minor Marine S<strong>and</strong> Sheets)<br />

These facies generally are unfossiliferous, very fine to fine <strong>and</strong><br />

less commonly medium-grained quartz–peloid grainstone com-<br />

?<br />

?<br />

?<br />

C11<br />

C10<br />

C9<br />

C8<br />

C7<br />

C6<br />

C5<br />

C4<br />

C3<br />

C2<br />

C1<br />

MERAMECIAN<br />

SEQUENCES


20<br />

Lithology<br />

& Depositional<br />

Environment<br />

Caliche, Breccia <strong>and</strong><br />

Erosional Contacts<br />

(Subaerial)<br />

Color Paleosols yellow to<br />

brown<br />

Depositional<br />

Texture <strong>and</strong><br />

Grain Types<br />

Sedimentary<br />

Structures<br />

Paleosols have<br />

cryptocrystalline <strong>and</strong><br />

fibrous calcite crusts <strong>and</strong><br />

fracture fills; rooted<br />

fabrics common, <strong>and</strong><br />

patches of caliche-coated<br />

peloids <strong>and</strong> pisolites.<br />

Variably silicified. May<br />

be brecciated.<br />

Paleosols have<br />

undulating lamination,<br />

rootlet casts, <strong>and</strong><br />

incipient brecciation.<br />

Erosional contacts are<br />

sharp, planar to<br />

scalloped surfaces on the<br />

underlying unit.<br />

Fossils Rootlet casts in<br />

paleosols.<br />

posed of well rounded peloids, rare ooids, skeletal fragments,<br />

<strong>and</strong> finer-grained subangular quartz concentrated at bases of<br />

laminae (Table 2; Figs. 6, 7). They are cross-stratified, with foresets<br />

of extremely planar to curved, paper-thin, millimeter-scale laminae,<br />

that commonly are inversely graded (from very fine to fine<br />

s<strong>and</strong>) <strong>and</strong> form a characteristic pinstripe pattern on the outcrop<br />

or in core. Where inverse grading is absent, laminae form couplets<br />

of alternating very fine <strong>and</strong> fine s<strong>and</strong>. These facies commonly<br />

are interbedded with red beds, lagoonal facies, <strong>and</strong> ooidshoal<br />

facies.<br />

These facies are interpreted as coastal eolianites on the basis<br />

of abundant quartz–peloid grainstone lithology, the well rounded<br />

<strong>carbonate</strong> grain shapes, planar to curved foresets with inverse<br />

grading, <strong>and</strong> very fine to fine s<strong>and</strong> couplets, pinstripe lamination,<br />

<strong>and</strong> lack of in situ fossils or burrows (Merkely, 1991; Hunter, 1993;<br />

Dodd et al., 1993; Smith, 1996). Some of the coarser-grained, lowangle<br />

cross-laminated units may be beach facies.<br />

Fine-Grained Lime Wackestone–Mudstone<br />

(Tidal Flat <strong>and</strong> Lagoon)<br />

This facies occurs in massive to laminated units 1 to 8 meters<br />

thick. They include thin to medium-bedded, light gray, finegrained<br />

lime wackestone–mudstone <strong>and</strong> pellet packstone, with a<br />

generally restricted assemblage of gastropods, ostracodes, <strong>and</strong><br />

rare small fragments of echinoderms <strong>and</strong> brachiopods (Table 2;<br />

Figs. 6, 7). Bedding planes locally contain abundant small bur-<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

TABLE 1.—Mississippian Paleosols <strong>and</strong> Siliciclastic Facies.<br />

Siliciclastic paleosols<br />

<strong>and</strong> red-beds<br />

(Subaerial to Marginal<br />

Marine)<br />

Red, maroon <strong>and</strong> green<br />

mottled<br />

Terrigenous mudrocks<br />

<strong>and</strong> siltstones.<br />

Redbeds massive, to<br />

laminated, rare current<strong>and</strong><br />

wave ripples on<br />

siltstone interbeds, rare<br />

mudcracks. Paleosols<br />

massive mudrocks with<br />

slickensides<br />

Root/ burrow traces in<br />

paleosols<br />

Gray shale<br />

(Low Energy Shallow<br />

Marine)<br />

Quartz s<strong>and</strong>stone/<br />

calcareous siltstones<br />

(Coastal to Shallow<br />

Marine)<br />

Dark gray to olive green White to light gray<br />

Terrigenous clay <strong>and</strong> silt Well-sorted fine to<br />

medium -grained shaly<br />

s<strong>and</strong>stones <strong>and</strong><br />

siltstones of quartz,<br />

lesser <strong>carbonate</strong> grains<br />

Poorly fissile to massive Cross-bedded to<br />

structureless; flaser,<br />

lenticular <strong>and</strong> wavy<br />

bedded locally<br />

Poorly fossiliferous to<br />

very fossiliferous.<br />

Mollusks, ostracods,<br />

some echinoderms<br />

brachiopods <strong>and</strong><br />

bryozoa; biota sparse<br />

<strong>and</strong> restricted in updip<br />

shale<br />

Rare<br />

rows <strong>and</strong> trails. In the east, some of these facies are microbially<br />

laminated dolomitic mudstones with centimeter layering, rare<br />

mudcracks, <strong>and</strong> few fossils. These laminated units are common<br />

downdip on the ramp in the eastern outcrop belt, but more<br />

massive units are widespread on the ramp (Table 2; Figs. 6, 7).<br />

The laminated facies are tidal-flat deposits developed downdip<br />

on the ramp during periods of emergence. The massive, finegrained<br />

<strong>carbonate</strong>s were formed in low-energy lagoonal settings<br />

on the ramp interior <strong>and</strong> between grainstone shoals on the ramp.<br />

Lack of open-marine fossils, <strong>and</strong> association with shallow-marine<br />

facies <strong>and</strong> exposure surfaces in updip areas, indicate a<br />

relatively restricted, possibly periodically hypersaline lagoonal<br />

setting.<br />

Peloid <strong>and</strong> Ooid Grainstone (Shoals <strong>and</strong> Tidal Bars)<br />

These are white, cross-bedded, well-sorted grainstones composed<br />

of medium-s<strong>and</strong>-size ooids, peloids, intraclasts, <strong>and</strong><br />

abraded skeletal fragments of echinoderms, bryozoa, brachiopods,<br />

<strong>and</strong> mollusks. They form units from 1 to 12 meters thick<br />

(Table 2; Figs. 6, 7). In the subsurface, the oolites form diporiented<br />

northwest-to-southeast trending oolitic s<strong>and</strong> ridges over<br />

30 km long <strong>and</strong> 1.4 km wide along the ramp margin (Kelleher <strong>and</strong><br />

Smosna, 1993; Smosna <strong>and</strong> Koehler, 1993). Both ooid <strong>and</strong> skeletal<br />

grainstones locally contain rare, small bryozoan mounds.<br />

These peloid–oolitic facies (Table 2) formed on high-energy<br />

tide-influenced <strong>and</strong> lesser wave-influenced ooid shoals in water


Facies<br />

& Depositional<br />

Environment.<br />

Quartz Peloid<br />

Grainstone<br />

(Coastal Eolianite,<br />

Minor Marine S<strong>and</strong><br />

Sheets)<br />

depths from the intertidal zone down to 10 meters (Kelleher <strong>and</strong><br />

Smosna, 1993; Keith <strong>and</strong> Zuppman, 1990).<br />

Skeletal Grainstone–Packstone<br />

(Ramp Margin <strong>and</strong> Moderate-Energy Lagoon)<br />

Because they are difficult to differentiate in well cuttings, all<br />

the coarse-grained, skeletal grainstones, skeletal packstones, <strong>and</strong><br />

grain-rich wackestones were lumped together. In outcrop, these<br />

are medium- to thick-bedded, <strong>and</strong> cross-bedded to structureless.<br />

They are medium- to coarse-grained skeletal units ranging from<br />

clean grainstones to muddy packstone <strong>and</strong> lesser wackestone.<br />

They consist of whole <strong>and</strong> fragmented crinoid, bryozoan, brachiopod,<br />

<strong>and</strong> lesser mollusk remains (Table 2). Some skeletal<br />

beds contain abundant whole columnals <strong>and</strong> calices. The skeletal<br />

grainstone units form extensive sheets seaward of oolitic<br />

grainstone shoals along the ramp margin in units up to 15 meters<br />

thick, which interfinger with wackestone <strong>and</strong> deep-water laminated<br />

marls (Figs. 6, 7). The more poorly washed facies form<br />

relatively extensive units on the ramp, interfingering laterally<br />

with <strong>and</strong> surrounded by poorly fossiliferous lime wackestone–<br />

mudstone (Figs. 6, 7).<br />

These facies include high-energy skeletal banks <strong>and</strong> sheets<br />

deposited in wave-agitated <strong>and</strong> tidal-current settings between<br />

<strong>and</strong> seaward of ooid shoals along the ramp margin, in water<br />

depths from a few meters to perhaps 20 meters. They also formed<br />

extensive units within the lagoon, in areas colonized by abundant<br />

crinoids <strong>and</strong> other skeletal organisms, <strong>and</strong> were subjected to low<br />

to moderate wave reworking.<br />

Argillaceous Skeletal Wackestone<br />

These are medium- to thick-bedded in outcrop <strong>and</strong> are dark<br />

gray shaly skeletal wackestones with some packstone interbeds.<br />

They are composed of medium- to very coarse-grained echinoderm,<br />

brachiopod, <strong>and</strong> bryozoan debris, with interstitial finely<br />

fragmented skeletal debris, lime mud, <strong>and</strong> argillaceous seams<br />

(Table 2). Some wackestone units have decimeter beds of skeletal<br />

grainstone–packstone. This facies locally contains very small<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

TABLE 2. —Mississippian Carbonate Facies.<br />

Fine-Grained Lime<br />

Wackestone/ Mudstone,<br />

Minor Lime Packstone<br />

(Tidal Flat <strong>and</strong> Low-<br />

Energy Lagoon)<br />

Color Dark-gray weathering Light gray. Tan to yellow<br />

weathering<br />

Grain Types And<br />

Depositional<br />

Texture<br />

Rounded <strong>and</strong> abraded<br />

peloids <strong>and</strong> some ooids,<br />

abraded skeletal<br />

fragments, <strong>and</strong><br />

subangular very fine to<br />

fine quartz up to 50%.<br />

Grainstone fabrics.<br />

Unfossiliferous to<br />

moderately fossiliferous<br />

fine wackestone <strong>and</strong><br />

mudstone <strong>and</strong> pellet<br />

packstone. Skeletal debris<br />

fine grained; abundant lime<br />

mud, dolomite mud, <strong>and</strong><br />

pellets; may contain quartz<br />

silt <strong>and</strong> clay. Locally<br />

cherty.<br />

Sedimentary<br />

Structures<br />

Wedge sets of planner<br />

<strong>and</strong> tangential crossbedding,<br />

with pinstripe<br />

lamination, <strong>and</strong> inverse<br />

grading, dips up to 20<br />

degrees<br />

Biota Abraded, rounded<br />

skeletal fragments. No<br />

in situ biota or marine<br />

trace fossils.<br />

Dolomitic mudstones<br />

massive to microbially<br />

laminated (centimeter-thick<br />

laminae, micrograded,<br />

rarely mudcracked). Finegrained,<br />

skeletal-rich<br />

wackestone-mudstone is<br />

typically massive.<br />

None to sparse; may have<br />

mollusks, small crinoid<br />

columnals, ostracods. May<br />

have trace-fossil horizons,<br />

small oncolites, rare corals,<br />

<strong>and</strong> brachiopods.<br />

Peloid <strong>and</strong> Ooid<br />

Grainstone<br />

(High-Energy<br />

Shoal)<br />

Skeletal Grainstone/<br />

Packstone<br />

(Ramp Margin <strong>and</strong><br />

Lagoonal Skeletal<br />

Sheets <strong>and</strong> Shoals)<br />

Argillaceous Skeletal<br />

Wackestone<br />

(Deep Subtidal)<br />

21<br />

bryozoan <strong>reefs</strong>, especially in sequences C10 <strong>and</strong> C11. Argillaceous<br />

packstone–wackestone is most abundant along the ramp<br />

margin (Figs. 6, 7). They formed below fair-weather wave base<br />

<strong>and</strong> above storm wave base on the open ramp in water depths of<br />

20 to 60 m by comparison with the Persian Gulf (Purser <strong>and</strong><br />

Seibold, 1973). However, these depths to maximum storm wave<br />

base could have been shallower in the Appalachian Basin, given<br />

that dominant wind directions in the Persian Gulf are axial to the<br />

basin <strong>and</strong> thus have maximum fetch, whereas dominant winds in<br />

the Appalachian Basin blew across the basin (Golonka et al.,<br />

1994). Siliciclastic muds may have washed out onto the ramp<br />

following floods or storms, or they may mark brief cessations in<br />

<strong>carbonate</strong> deposition during short periods of more humid climate<br />

<strong>and</strong> higher siliciclastic influx.<br />

Laminated Shaly Lime Mudstone (Ramp Slop to Basin)<br />

These are unfossiliferous, dark gray, evenly thin-laminated to<br />

massive, commonly silty or argillaceous lime-mudstone units up<br />

to 100 meters thick along the ramp slope (Fig. 6, section 11; Table 2).<br />

They consist of fine, silt-size <strong>carbonate</strong> <strong>and</strong> quartz, finely comminuted<br />

skeletal debris, <strong>and</strong> a matrix of argillaceous lime mudstone.<br />

The shaly lime mudstones are deep-water ramp-slope to<br />

basin-floor facies, which lack shallow-water sedimentary structures<br />

<strong>and</strong> biota. Fine <strong>carbonate</strong> sediment was brought onto the<br />

slope perhaps by storms <strong>and</strong> tidal currents <strong>and</strong> mixed with l<strong>and</strong>derived<br />

terrigenous silts <strong>and</strong> clays. Lack of burrowing suggests<br />

deposition in anoxic waters below the photic zone in depths of 60<br />

to over 100 m; the photic zone in the basin perhaps extended<br />

down to only to 30 m, given the turbid basin waters, whereas<br />

wave base perhaps reached down to more than 30 m but less than<br />

60 m, by comparison with the present Persian Gulf, a similarsized<br />

<strong>and</strong> similarly situated forel<strong>and</strong> basin (Purser <strong>and</strong> Seibold,<br />

1973).<br />

SEQUENCE STRATIGRAPHY<br />

Laminated Shaly<br />

Lime Mudstone<br />

(Ramp slope-<br />

Basin)<br />

Light gray to white Light to medium gray Medium gray to dark gray Typically dark gray<br />

Well sorted,<br />

rounded, medium to<br />

coarse grainstone.of<br />

s<strong>and</strong>-size ooids,<br />

peloids, lesser<br />

skeletal fragments,<br />

<strong>and</strong> minor<br />

intraclasts.<br />

Cross-bedded to<br />

massive<br />

Crinoid , brachiopod,<br />

bryozoan, <strong>and</strong><br />

mollusk fragments,<br />

forams<br />

Variably fragmented<br />

echinoderms,<br />

brachiopod, bryozoa,<br />

mollusks, <strong>and</strong> rare ooids.<br />

Mud-free grainstones to<br />

grain-rich packstones<br />

<strong>and</strong> minor wackestones,<br />

some with argillaceous<br />

seams.<br />

May be cross-bedded,<br />

thin- to medium-bedded<br />

or structureless<br />

Abundant echinoderms,<br />

brachiopods, <strong>and</strong><br />

bryozoas; <strong>and</strong> lesser<br />

mollusks.<br />

Echinoderms,<br />

brachiopods, bryozoa, <strong>and</strong><br />

lesser mollusks. Include<br />

wackestone <strong>and</strong> lesser<br />

packstone; abundant lime<br />

mud; terrigenous clay<br />

disseminated or in seams<br />

<strong>and</strong> stringers<br />

Massive, thick to thin<br />

bedded, structureless;<br />

may have graded storm<br />

beds<br />

Abundant echinoderm,<br />

common brachiopods <strong>and</strong><br />

bryozoas, rare mollusks<br />

Carbonate <strong>and</strong><br />

silisiclastic clay <strong>and</strong><br />

silt<br />

Millimeter to<br />

centimeter even<br />

lamination to<br />

structureless<br />

None<br />

The sequence stratigraphic terms used are after Sarg (1988),<br />

Mitchum <strong>and</strong> Van Wagoner (1991), <strong>and</strong> Weber et al. (1995). Facies


22<br />

<strong>and</strong> depositional sequences are shown in Figure 6A (extending<br />

northeast to southwest down the axis of the foredeep <strong>and</strong> into the<br />

proximal forel<strong>and</strong> basin, with sections arranged from thinnest to<br />

thickest) <strong>and</strong> in Figure 7A (oriented east–west at right angles to<br />

the forel<strong>and</strong> basin, <strong>and</strong> extending across the distal forel<strong>and</strong>, <strong>and</strong><br />

just beyond the hinge into the margin of the proximal forel<strong>and</strong>).<br />

Depositional sequences <strong>and</strong> their component facies were traceable<br />

from the outcrop sections into the subsurface using well<br />

cuttings <strong>and</strong> wireline logs, especially gamma-ray logs.<br />

Sequences were recognized on the basis of major l<strong>and</strong>ward<br />

<strong>and</strong> basinward shifts in diagnostic facies belts such as eolianites,<br />

red beds, siliciclastic s<strong>and</strong>s <strong>and</strong>/or calcareous siltstones, <strong>and</strong> lime<br />

grainstones. Sequence boundaries were picked at erosional disconformities<br />

(commonly evident in outcrop sections) on top of<br />

shallowing-upward <strong>carbonate</strong> units, <strong>and</strong> below lowst<strong>and</strong> or<br />

transgressive siliciclastic red beds, s<strong>and</strong>stones, <strong>and</strong> shale (evident<br />

in both subsurface <strong>and</strong> outcrop). Along the ramp slope,<br />

sequence boundaries were placed beneath the shallowest-water<br />

facies within the dominantly deep-water successions. In wireline<br />

logs, many sequence boundaries (<strong>and</strong> some parasequence boundaries)<br />

are associated with significant gamma-ray kicks due to<br />

presence of terrigenous mudrocks, shale, or argillaceous dolomites.<br />

This was important in sections containing parasequences,<br />

because it allowed parasequences to be grouped into sequences<br />

on the basis of the characteristic upward-decreasing (as the<br />

limestones clean upward) to upward-increasing gamma-ray signature<br />

(as the <strong>carbonate</strong>s become more siliciclastic-rich). Because<br />

it was not clear using the subsurface data whether siliciclastic<br />

units were late highst<strong>and</strong>, lowst<strong>and</strong>, or early transgressive, they<br />

have been arbitrarily referred to as lowst<strong>and</strong> if on the ramp slope<br />

<strong>and</strong> margin, <strong>and</strong> lowst<strong>and</strong> to early transgressive if updip. Where<br />

possible, maximum flooding surfaces were picked below the<br />

deepest marine facies within the sequence, but the subsurface<br />

data commonly prevented them being traced regionally; thus it<br />

was possible to separate the transgressive <strong>and</strong> highst<strong>and</strong> systems<br />

tracts only locally on the ramp. Some of the maximum flooding<br />

surfaces are associated with high gamma-ray values, especially<br />

where the <strong>carbonate</strong>s are shaly, but many in pure <strong>carbonate</strong><br />

sections are not.<br />

Fourth-order sequences (roughly between 0.1 <strong>and</strong> 0.5 My<br />

duration) are the dominant units mapped in the study interval in<br />

the Appalachian Basin <strong>and</strong> the Illinois Basin (Smith <strong>and</strong> Read,<br />

1999, 2001; Al-Tawil <strong>and</strong> Read, this volume). Third-order or<br />

composite sequences (Weber et al., 1995) are far less well developed,<br />

being masked by the profound facies shifts within the<br />

dominant fourth-order sequences. Individual fourth-order sequences<br />

<strong>and</strong> component facies are shown in Figures 6B <strong>and</strong> 7B,<br />

constructed by rehanging the sections from the top of each<br />

sequence. Major biostratigraphic markers include Acrocyathus in<br />

the Hillsdale Formation, the last appearance of Platycrinites penicillus<br />

at the top of the Denmar Formation, the first appearance of<br />

Talarocrinus in the Pickaway Formation, as well as the conodont<br />

zones (Fig. 3). Individual fourth-order sequences were below the<br />

resolution of the biostratigraphic data, which typically span<br />

several sequences, but they do form valuable biostratigraphic<br />

markers within the succession.<br />

The sparse radiometric data <strong>and</strong> their error bars, <strong>and</strong> the<br />

difficulties in intercontinental correlation between Mississippian<br />

brachiopod, conodont, <strong>and</strong> foraminiferal zones, make absolute<br />

age determination of the study interval difficult. The St. Louis–<br />

Hillsdale interval is probably about 5 My in duration (Ross <strong>and</strong><br />

Ross, 1987); sequences C1 to C11 (the Denmar to Glenray interval)<br />

are about 3 to 8 My (S<strong>and</strong>o, 1984; Baxter <strong>and</strong> Brenckler, 1982;<br />

Roberts et al., 1995; Smith, 1996). Assigning duration of 3 to 8 My<br />

for the three third-order sequences in the Denmar to Glenray<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

interval gives a duration of 1 to 2.5 My for each of the composite<br />

sequences <strong>and</strong> 280 ky to an extreme upper limit of 720 ky for each<br />

of the 11 fourth-order sequences, with the most likely durations<br />

being 300 to 400 ky. (Smith, 1996). This encompasses the estimated<br />

average duration of 375 ky for the time-equivalent fourthorder<br />

sequences in the Illinois Basin (Smith, 1996) <strong>and</strong> the 400 ky<br />

durations estimated for the overlying Mauch Chunk siliciclastic<br />

fourth-order sequences (Miller <strong>and</strong> Eriksson, 1999).<br />

Fourth-Order Depositional Sequences<br />

Meramecian Sequences.—<br />

These Meramecian sequences (Figs. 6, 7) form the basal part of<br />

the study interval <strong>and</strong> include the Little Valley <strong>and</strong> Hillsdale<br />

formations. They unconformably (?) overlie the Price–Maccrady<br />

red beds, resting on a widespread transgressive s<strong>and</strong>stone in<br />

many places. The Meramecian sequences thicken from 0 to over<br />

270 m across the hinges into the basin (Fig. 6, section 11; Bartlett<br />

<strong>and</strong> Webb, 1971). Two to four Meramecian sequences may be<br />

present on the ramp. The sequence boundaries are placed beneath<br />

widespread lowst<strong>and</strong> to early transgressive s<strong>and</strong>stone–<br />

calcareous siltstone on the ramp to the west (Fig. 7), which appear<br />

to pinch out to the north, where two sequences may make up the<br />

interval (sequences not plotted individually in Fig. 6B). The<br />

transgressive to highst<strong>and</strong> parts of the sequences consist mainly<br />

of commonly cherty, fine skeletal wackestone–mudstone, grading<br />

basinward into shaly lime wackestone or lime mudstone. In<br />

the thick, proximal forel<strong>and</strong> basin sections, several sequences<br />

may be present (not shown in Fig. 6). Lower sequences here<br />

(mapped as Little Valley Formation) consist of quartz s<strong>and</strong>stone<br />

overlain by dark gray, nearshore argillaceous skeletal wackestone–<br />

mudstone, whereas the overlying Hillsdale sequences consist of<br />

oncolitic skeletal wackestone–mudstone capped by skeletal<br />

grainstone (Huggins, 1983).<br />

Chesterian Sequences.—<br />

Sequence C1.—Sequence C1 is the basal sequence of the Denmar<br />

Limestone (Figs. 6, 7). It generally overlies the Upper Meramecian<br />

Hillsdale Limestone <strong>and</strong> is 0 to 55 m thick (Fig. 6). The basal<br />

sequence boundary is placed beneath widespread basal, thin<br />

siliciclastic-prone units. Sequence C1 has a basal lowst<strong>and</strong> unit<br />

of local red beds that grades seaward into quartz s<strong>and</strong>stone–<br />

siltstone a few meters thick along the ramp margin <strong>and</strong> ramp<br />

slope (Fig. 6, sections 7 to 9 <strong>and</strong> 11; Fig. 7, sections 25, 26). These<br />

grade updip onto the ramp interior into lowst<strong>and</strong>–transgressive<br />

s<strong>and</strong>stone–calcareous siltstone, shale, <strong>and</strong> quartz–peloid<br />

grainstones (Fig. 6; sections 3 to 6; Fig. 7, sections 19 to 26) that<br />

may be overlain by possible transgressive <strong>carbonate</strong>s in the<br />

middle of the sequence in the west (Fig. 7). The transgressive<br />

systems tract along the ramp slope consists of up to 20 meters of<br />

skeletal wackestone–mudstone grading up into skeletal grainstone<br />

(Fig. 6, section 11). The highst<strong>and</strong> systems tract downdip consists<br />

of deeper-water, laminated mudstones <strong>and</strong> skeletal wackestone<br />

that passes upsection into skeletal grainstone. Updip the highst<strong>and</strong><br />

systems tract consists of thin, ramp-margin skeletal–ooid<br />

grainstones that grade updip into lime wackestone–mudstone<br />

<strong>and</strong> discrete ooid–peloid grainstone shoal deposits (Fig. 6, sections<br />

1 to 11; Fig. 7, sections 19 to 26).<br />

Sequence C2.—Sequence C2, the next Denmar sequence, is 0 to<br />

75 m thick (Figs. 6, 7); it thickens markedly between sections 6 <strong>and</strong><br />

7 <strong>and</strong> sections 10 <strong>and</strong> 11 (Fig. 6). The sequence boundary is placed<br />

beneath a thin quartz s<strong>and</strong>stone on the ramp margin <strong>and</strong> slope,


<strong>and</strong> updip beneath widespread siltstone to the west (Fig. 7) <strong>and</strong><br />

beneath eolianites updip to the north (Fig. 6). It has a lowst<strong>and</strong><br />

quartz s<strong>and</strong>stone a few meters thick along the ramp margin <strong>and</strong><br />

ramp slope (Fig. 6, section 11; Fig. 7, sections 24 to 26). The<br />

transgressive systems tract along the ramp slope consists of<br />

skeletal wackestone (Fig. 6, section 11; Fig. 7, section 26), <strong>and</strong><br />

within a broad sag on the ramp margin (Fig. 6, sections 6 to 10),<br />

it probably includes the lower third of the sequence, consisting of<br />

quartz peloid <strong>and</strong> ooid grainstone <strong>and</strong> lime wackestone–mudstone<br />

beneath the lowest skeletal grainstones. The lowst<strong>and</strong> to<br />

transgressive systems tract updip includes thin quartz–peloid<br />

eolianites, <strong>and</strong> the lower portion of ooid grainstones in the<br />

northeast (Fig. 6, sections 1 to 5); to the west, the lowst<strong>and</strong> to<br />

transgressive tract consists of the regional calcareous siltstones<br />

(Fig. 7). The highst<strong>and</strong> systems tract along the ramp slope consists<br />

of deep-water shaly lime mudstone overlain by skeletal<br />

wackestone (Fig. 6, section 11; Fig. 7, section 26). Farther updip it<br />

consists of ramp-margin skeletal <strong>and</strong> minor ooid grainstones<br />

(Fig. 6, sections 9, 10; Fig. 7, section 25) that interfinger with rampinterior<br />

lime wackestone–mudstone containing multilateral ooid<br />

grainstone units (Figs. 6, 7). The vertical repetition of facies<br />

suggests that two to three parasequences may make up the<br />

sequence between sections 6 to 10 (Fig. 6).<br />

Sequence C3.—This next Denmar sequence is 0 to 40 m thick<br />

(Figs. 6, 7). Its basal sequence boundary is placed beneath thin<br />

siliciclastic units or eolianite (Figs. 6, 7). The lowst<strong>and</strong> systems<br />

tract is a thin quartz s<strong>and</strong>stone on the ramp slope (Fig. 6, section<br />

11). The transgressive systems tract is a wedge of skeletal wackestone<br />

overlain by skeletal grainstone along the ramp slope (Fig. 6,<br />

section 11). Updip the transgressive systems tract may include<br />

the lower half of the sequence, <strong>and</strong> is composed of thin, basal<br />

siliciclastic <strong>and</strong> eolianite units <strong>and</strong> the overlying lower parts of<br />

the discontinuous ooid grainstone <strong>and</strong> lime wackestone–mudstone<br />

units (Figs. 6, 7). The highst<strong>and</strong> systems tract constitutes the<br />

upper part of the sequence <strong>and</strong> consists of a patchwork of skeletal,<br />

oolitic, <strong>and</strong> lime wackestone–mudstone facies (Figs. 6, 7).<br />

Sequence C4.—Sequence C4, the top Denmar sequence, ranges<br />

from 5 to 80 m thick with marked thickening between sections 10<br />

<strong>and</strong> 11 (Fig. 6) <strong>and</strong> sections 24 to 26 (Fig. 7). On the ramp margin,<br />

the sequence boundary is a double disconformity in outcrop (Fig.<br />

6, section 10), <strong>and</strong> on the ramp slope the sequence boundary is<br />

beneath lowst<strong>and</strong> quartz s<strong>and</strong>stone several meters thick (Fig. 6,<br />

section 11) or calcareous siltstone (Fig. 7, sections 24 to 26). The<br />

sequence boundary on the ramp interior is placed beneath discontinuous<br />

lowst<strong>and</strong> eolianites (Fig. 6, sections 3, 7) or thin basal<br />

siliciclastic or quartz–peloid grainstone units (Fig. 7). The transgressive<br />

systems tract is absent on the slope, but on the ramp it<br />

may include thin local units of ooid grainstone <strong>and</strong> quartz–peloid<br />

grainstone in the north <strong>and</strong> east (Fig. 6, sections 3, 7, 10); in the<br />

west, the lowst<strong>and</strong> to transgressive systems tract consists of an<br />

extensive basal siliciclastic sheet (Fig. 7, sections 19 to 23). The<br />

highst<strong>and</strong> systems tract makes up the bulk of the sequence, <strong>and</strong><br />

consists of ramp-margin skeletal grainstone–packstone <strong>and</strong> ooid–<br />

peloid grainstone, <strong>and</strong> ramp-interior, discontinuous units of lime<br />

wackestone–mudstone that separate laterally, local ooid–peloid<br />

shoal deposits (Figs. 6, 7). On the ramp slope, the highst<strong>and</strong> tract<br />

consists of deeper-water shaly mudstones (Fig. 6, section 11).<br />

Sequence C5.—Sequence C5, 0 to 25 m thick, includes the<br />

lower Taggard red beds (Figs. 6, 7). It shows a considerable<br />

basinward shift in red siliciclastic facies compared with the<br />

underlying sequences. The sequence boundary is beneath widespread<br />

siliciclastics <strong>and</strong> local quartz–peloid eolianites. The<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

23<br />

lowst<strong>and</strong> to early transgressive systems tract consists of red<br />

beds updip in the northeast (Fig. 6) <strong>and</strong> on the ramp margin in<br />

the west (Fig. 7, section 26); thin, local eolianites <strong>and</strong> s<strong>and</strong>stone–<br />

siltstone on ramp in the west (Fig. 7, sections 19 to 22); <strong>and</strong> up<br />

to 15 m of s<strong>and</strong>stone–calcareous siltstone <strong>and</strong> marine shale on<br />

the ramp margin <strong>and</strong> slope (Fig. 6, sections 8, 9; Fig. 7, sections<br />

20 to 25). No transgressive systems tract is recognized in the<br />

north <strong>and</strong> east (Fig. 6). In the west, however, the transgressive<br />

tract may include the lower half of the <strong>carbonate</strong> interval,<br />

composed of basal units of ooid <strong>and</strong> peloid grainstone <strong>and</strong> the<br />

overlying lime wackestone–mudstones, if these are shallow<br />

lagoonal facies (Fig. 7, sections 17 to 22). The highst<strong>and</strong> tract on<br />

the shallow ramp is dominated by lime wackestone–mudstone,<br />

with ramp-margin grainstone units, <strong>and</strong> localized updip ooid<br />

<strong>and</strong> skeletal grainstone bodies (Figs. 6, 7); the highst<strong>and</strong> systems<br />

tract on the slope consists of basinal shaly lime mudstone<br />

grading up into skeletal wackestone (Fig. 6, section 11).<br />

Sequence C6.—Sequence C6 is 0 to 40 m thick (Figs. 6, 7) <strong>and</strong><br />

thickens rapidly between sections 5 <strong>and</strong> 6 (Fig. 6). It includes the<br />

upper Taggard red beds <strong>and</strong> the lower Pickaway Limestone. The<br />

sequence boundary is placed below extensive siliciclastics (Fig. 6)<br />

or beneath widespread quartz–peloid eolianites (Fig. 7). The<br />

lowst<strong>and</strong> to early transgressive deposits consist of regional red<br />

beds up to 10 m thick in the north <strong>and</strong> east (Fig. 6) or regional<br />

eolianites to the west (Fig. 7). Downdip these grade into up to 30<br />

m of lowst<strong>and</strong> marine s<strong>and</strong>stone locally split by a thin shale–<br />

calcareous siltstone (Fig. 6, sections 9 to 11; Fig. 7, sections 20 to<br />

26). The transgressive systems tract on the ramp slope consists of<br />

a 15-m-thick succession of skeletal grainstone capped by a thin<br />

ooid grainstone (Fig. 6, section 11), which appears to grade updip<br />

into peloid grainstone (Fig. 7, section 26). On the ramp, it is<br />

difficult to separate the transgressive systems tract from the<br />

highst<strong>and</strong> systems tract. These combined systems tracts are 0 to<br />

35 m thick. Along the ramp margin in the west, probable highst<strong>and</strong><br />

deposits include a downdip ooid grainstone unit grading updip<br />

into skeletal wackestone <strong>and</strong> then into ramp-margin skeletal<br />

grainstone–packstone (Fig. 7, sections 23 to 26); these likely grade<br />

downdip into skeletal grainstone, skeletal wackestones, <strong>and</strong> then<br />

basinal lime mudstone (Fig. 6, section 11). On the ramp interior in<br />

the northeast, the transgressive to highst<strong>and</strong> systems tract is<br />

dominated by widespread units of lime wackestone–mudstone<br />

(Figs. 6, 7) with eolianites in the far west. Skeletal grainstone–<br />

packstone <strong>and</strong> minor ooid grainstones also are localized within<br />

the upper third of the sequence in a broad sag on the middle ramp<br />

in the northeast (Fig. 6, sections 6, 7).<br />

Sequence C7.—Sequence C7 includes the upper Pickaway<br />

Limestone. It is 0 to 50 m thick, thickening between sections 3 <strong>and</strong><br />

4 (Fig. 6) <strong>and</strong> between sections 22 to 26 (Fig. 7). The C7 sequence<br />

boundary on the ramp <strong>and</strong> ramp margin is placed below lowst<strong>and</strong><br />

to early transgressive s<strong>and</strong>stone, shale, calcareous siltstone, minor<br />

red beds, <strong>and</strong> quartz–peloid eolianites in the east <strong>and</strong> northeast<br />

(Fig. 6, sections 5 to 9; Fig. 7, sections 18 to 26), <strong>and</strong> beneath<br />

skeletal wackestone on the ramp slope (Fig. 6, section 11). The<br />

transgressive systems tract over much of the ramp in the northeast<br />

contains skeletal grainstone interfingering laterally with<br />

lime wackestone–mudstone, <strong>and</strong> is capped downdip by a thin<br />

shale <strong>and</strong> s<strong>and</strong>stone <strong>and</strong> updip by s<strong>and</strong>stone <strong>and</strong> ooid grainstone<br />

(Fig. 6, sections 3 to 9); the transgressive succession is not clearly<br />

defined in the west but may include a wedge of skeletal wackestone<br />

overlain by calcareous siltstone–shale along the ramp margin<br />

(Fig. 7, sections 24 to 26). On the ramp slope, the transgressive<br />

systems tract is a skeletal wackestone (Fig. 6, section 11). The<br />

highst<strong>and</strong> tract along the ramp slope is a deep-water shaly lime


24<br />

mudstone (Fig. 6, section 11). On the ramp, the highst<strong>and</strong> systems<br />

tract in the east contains two belts of skeletal–ooid grainstone, one<br />

along the ramp margin (Fig. 6, section 9) <strong>and</strong> the other updip to<br />

the north flanking a significant downwarp (Fig. 6, sections 4 to 6);<br />

these updip grainstones are bordered both updip <strong>and</strong> downdip<br />

by lime wackestone–mudstone. In the west, the highst<strong>and</strong> is a<br />

simple facies succession of skeletal grainstone <strong>and</strong> skeletal<br />

wackestone that passes updip into ooid limestone <strong>and</strong> then into<br />

lime wackestone–mudstone (Fig. 7). Three or more parasequences<br />

may be present in the sequence.<br />

Sequence C8.—Sequence C8 consists of the lower Union Limestone<br />

<strong>and</strong> is 0 to 55 m thick (Figs. 6, 7). It thickens between sections<br />

3 <strong>and</strong> 7 (Fig. 6). The C8 sequence boundary is placed beneath<br />

lowst<strong>and</strong> to early transgressive deeper-water skeletal wackestones<br />

along the lower ramp slope (Fig. 6, section 11), <strong>and</strong> beneath<br />

lowst<strong>and</strong> to transgressive s<strong>and</strong>stone, calcareous siltstone, <strong>and</strong><br />

rare red beds on the upper ramp slope <strong>and</strong> ramp margin (Fig. 6,<br />

section 9; Fig. 7, sections 22 to 26). The lowst<strong>and</strong> to early transgressive<br />

systems tract on the ramp interior may include basal,<br />

thin, patchily developed shales, peloid grainstone, <strong>and</strong> quartz–<br />

peloid eolianites (Fig. 6, 7). The transgressive systems tract on the<br />

ramp margin in the east makes up the lower one third of the<br />

sequence, <strong>and</strong> includes skeletal wackestone (Fig. 6, section 11)<br />

grading upsection <strong>and</strong> updip into skeletal grainstone <strong>and</strong> lesser<br />

ooid grainstone (Fig. 6, sections 7 to 9) <strong>and</strong> thin siliciclastic units<br />

<strong>and</strong> quartz–peloid eolianite (Fig. 6, sections 3 to 5; Fig. 7, sections<br />

18 to 25). The highst<strong>and</strong> systems tract is thick <strong>and</strong> dominated by<br />

an upward-shallowing succession of deeper-water mudstone to<br />

skeletal wackestone along the ramp slope (Fig. 6, section 11). On<br />

the ramp margin <strong>and</strong> ramp interior, the highst<strong>and</strong> systems tract<br />

includes at least two shallowing-upwards successions of skeletal<br />

wackestone up into skeletal grainstone that is bordered updip by<br />

locally thick ooid grainstone (Fig. 6, sections 8, 9; Fig. 7, sections<br />

23 to 26); these grade updip into lime wackestone–mudstone with<br />

local units of ooid grainstone (Fig. 6), <strong>and</strong> in the west, some thick<br />

eolianites (Fig. 7, sections 12, 19) separated laterally by lime<br />

wackestone–mudstone <strong>and</strong> containing a local shale in the middle<br />

of the succession (Fig. 6; Fig. 7, sections 18, 19).<br />

Sequence C9.—Sequence C9 is 5 to 25 m thick, <strong>and</strong> includes the<br />

upper Union Limestone (Figs. 6, 7). Updip in outcrop, the basal<br />

sequence boundary is locally at the base of thin channeled s<strong>and</strong>stones<br />

<strong>and</strong> red beds (Fig. 6, sections 1, 2). Farther downdip, the<br />

sequence boundary is placed beneath a thin local shale <strong>and</strong> a<br />

widespread unit of calcareous siltstone <strong>and</strong> minor s<strong>and</strong>stone<br />

(Figs. 6, 7) <strong>and</strong> beneath a thin oolite on the ramp slope (Fig. 6,<br />

section 11). This thin ooid grainstone is the lowst<strong>and</strong> systems<br />

tract along the ramp slope. Updip, the lowst<strong>and</strong> to transgressive<br />

systems tract consists of calcareous siltstone <strong>and</strong> shale (Fig. 6,<br />

sections 5 to 9; Fig. 7, sections 15 to 26) <strong>and</strong> in the far north, the thin<br />

channeled s<strong>and</strong>stones capped by red beds (mid-Union red beds)<br />

(Fig. 6, sections 1, 2). The transgressive systems tract on the ramp<br />

slope is a thin skeletal wackestone above the lowst<strong>and</strong> oolite (Fig.<br />

6, section 11), but a distinct transgressive systems tract is not<br />

easily recognized updip. The highst<strong>and</strong> systems tract is an upward-shallowing<br />

succession of deeper-water shaly lime mudstone<br />

to skeletal wackestone on the lower ramp slope (Fig. 6,<br />

section 11; Fig. 7, section 26). These pass upslope to the north into<br />

extensive ramp-margin skeletal grainstone (Fig. 6, sections 7 to 9;<br />

Fig. 7, sections 24, 25). These ramp-margin grainstones pass<br />

l<strong>and</strong>ward into skeletal wackestone or lime wackestone–mudstone,<br />

<strong>and</strong> then into updip skeletal grainstones <strong>and</strong> ooid grainstone<br />

(Fig. 6, sections 2 to 5; Fig. 7, sections 20 to 22) <strong>and</strong> on the<br />

innermost ramp, lime wackestone–mudstone.<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

Sequence C10.—Sequence C10, 10 to 60 m thick, consists of the<br />

Greenville Shale <strong>and</strong> Alderson Limestone (Figs. 6, 7). The C10<br />

basal sequence boundary is placed beneath a widespread thin<br />

shale or beneath calcareous siltstone <strong>and</strong> s<strong>and</strong>stone (Figs. 6, 7).<br />

The lowst<strong>and</strong> systems tract along the ramp slope is a thin s<strong>and</strong>stone–siltstone<br />

(Fig. 6, section 11; Fig. 7, sections 24 to 26). The<br />

lowst<strong>and</strong> to transgressive systems tract farther updip probably<br />

includes the basal shale <strong>and</strong> the calcareous siltstone unit (Figs. 6,<br />

7). The highst<strong>and</strong> systems tract consists of skeletal wackestone<br />

downdip, passing updip into ramp-margin skeletal grainstone–<br />

packstone (Fig. 6, sections 7 to 9; Fig. 7, sections 25, 26). These<br />

grade updip to the northeast into thin ooid grainstone units <strong>and</strong><br />

widespread lime wackestone–mudstone (Fig. 6, sections 2 to 6) or<br />

in the west, into thick ooid–peloid grainstone <strong>and</strong> then into lime<br />

wackestone–mudstone <strong>and</strong> local units of skeletal grainstone–<br />

packstone (Fig. 7, sections 12 to 22).<br />

Sequence C11.—Sequence C11 includes the Lillydale Shale <strong>and</strong><br />

the Glenray Limestone <strong>and</strong> is 0 to over 100 m thick thickening<br />

rapidly across the hinge (Fig. 6, sections 9 to 11; Fig. 7, sections 21<br />

to 22). The C11 sequence boundary is placed beneath thin, locally<br />

developed quartz s<strong>and</strong>stone–siltstones (Fig. 6, sections 1, 6, 8; Fig.<br />

7, sections 17 to 20), or beneath thin shales, red beds, or eolianites<br />

(Fig. 7, sections 12 to 16), or at the erosional contact on C10<br />

<strong>carbonate</strong>s at section 2 (Fig. 6). In the west, the sequence boundary<br />

on the ramp may be locally incised, indicated by rapid<br />

thickening of the siltstone at the expense of underlying units (Fig.<br />

6, section 11; Fig. 7, section 18). The lowst<strong>and</strong> systems tract on the<br />

ramp slope is a cross-bedded lithoclastic s<strong>and</strong>stone (Fido S<strong>and</strong>stone),<br />

which appears to be incised into the underlying unit (Fig.<br />

6, section 11). The transgressive systems tract is not defined<br />

downdip, but updip the lowst<strong>and</strong> to transgressive systems tract<br />

consists of thin, local s<strong>and</strong>stone–siltstones (which thicken locally<br />

into the possible incised valley at section 18, Fig. 7), <strong>and</strong> an<br />

overlying thin regional oolitic <strong>and</strong> skeletal limestone (Figs. 6, 7).<br />

The highst<strong>and</strong> systems tract includes 10 to 40 m of fossiliferous<br />

marine shale (Lillydale Shale) that grades basinward into a rampmargin<br />

complex of skeletal wackestone, skeletal grainstone, <strong>and</strong><br />

interbedded s<strong>and</strong>stone–siltstone (Fig. 7, section 26) <strong>and</strong> then into<br />

basinal lime mudstone up to 80 m thick (Fig. 6, section 11). These<br />

are overlain by late highst<strong>and</strong>, thin, discontinuous skeletal limestone<br />

capped locally by ooid grainstone (Fig. 6, sections 1, 3, 6, 8,<br />

9, 11; Fig. 7, sections 22 to 23). The top sequence boundary of<br />

sequence C11 in outcrop is a paleosol updip <strong>and</strong>/or erosional<br />

surface updip (Fig. 6, sections 1, 3), but on the ramp slope (Fig. 6,<br />

section 11) the sequence boundary underlies a thin black shale<br />

capped by ooid grainstone that is conformable on C11 skeletal<br />

wackestone. Sequence C11 is overlain by highly progradational<br />

red <strong>and</strong> gray siliciclastics (Mauch Chunk Formation).<br />

Parasequences<br />

Fourth-order sequences typically have only one or two<br />

parasequences updip, but downdip, sequences commonly are<br />

composed of several parasequences. Detailed facies successions<br />

of parasequences <strong>and</strong> presence or absence of erosional disconformities<br />

capping units are best observed in the outcrop sections or<br />

core. No attempt has been made to show parasequences in<br />

Figures 6 <strong>and</strong> 7. The parasequences are difficult to recognize in<br />

the cuttings data in sections in which the parasequences are thin<br />

<strong>and</strong> of the same magnitude as the sample spacing of the cuttings.<br />

Siliciclastic-prone parasequences in updip areas commonly<br />

consist of marine s<strong>and</strong>stone, locally with a slightly channeled<br />

base, fining upward into siltstone <strong>and</strong> then into red mudrock.<br />

Others include thin ooid or peloid grainstone overlain by red or


gray siltstone <strong>and</strong> then red mudrock. Siliciclastic units commonly<br />

make up the lower two-thirds of parasequences at the bases of<br />

sequences. Disconformity-bounded parasequences (2 to 6 m thick)<br />

are recognizable even in thick successions of restricted lime<br />

wackestone–mudstone (as in C7) <strong>and</strong> are composed of very thin<br />

(or absent) basal grainstones overlain by wackestone–mudstone,<br />

in which layering increases in abundance <strong>and</strong> becomes thinner<br />

upward into the laminated caps, where present.<br />

Oolite-dominated sequences include asymmetrical, upwardshallowing<br />

units (3 to 10 m thick) bounded at the base or top by<br />

quartz–peloid eolianites, <strong>and</strong> are composed of skeletal wackestone<br />

(commonly absent), up into ooid grainstone (with rare channeled<br />

bases <strong>and</strong> local mounds), which may be capped by argillaceous<br />

dolomitic wackestone–mudstone or dolomitic mudstone. Erosional<br />

disconformities may occur on the top of the lime<br />

wackestone–mudstone units, or on the ooid grainstone units.<br />

Symmetrical upward-deepening to upward-shallowing oolitic<br />

parasequences contain basal lime wackestone–mudstone conformably<br />

overlain by the upward-shallowing succession.<br />

Parasequences in skeletal limestone–rich parts of sequences<br />

(sequences C8 to C10) commonly are 2 to 15 m thick, <strong>and</strong><br />

generally are bounded by erosional or sharp planar contacts.<br />

Such parasequences include asymmetrical, upward-shallowing<br />

parasequences composed of thick skeletal wackestone–<br />

packstone grading up into ooid grainstone (may be absent),<br />

capped by dolomitic wackestone–mudstone. Other skeletal limestone–dominated<br />

parasequences are symmetrical <strong>and</strong> contain<br />

basal, thin transgressive ooid or skeletal grainstone, or basal<br />

upward-deepening units of lime wackestone–mudstone,<br />

interbedded with thin grainstones. Parasequences that straddle<br />

the maximum flooding surfaces of sequences commonly are<br />

thick <strong>and</strong> symmetrical.<br />

Third-Order or Composite Sequences<br />

Third-order or composite sequences (Weber et al., 1995) are<br />

relatively weakly developed in the study interval compared to<br />

the regionally mappable fourth-order sequences. The third-order<br />

sequences (Figs. 6, 7) are bounded by red beds <strong>and</strong> their downdip<br />

marine s<strong>and</strong>stones updip. They thicken downdip into the<br />

foredeep, where the third-order sequences are bounded along the<br />

ramp margin <strong>and</strong> ramp slope by extensive lowst<strong>and</strong> siliciclastic<br />

units. The third-order sequences commonly contain two or more<br />

high-frequency sequences.<br />

Composite Sequence 1.—Composite sequence 1 is composed of<br />

the Meramecian Little Valley–Hillsdale interval. It extends from<br />

the top of the Price–Maccrady red beds, up to the red beds of<br />

sequence C1 <strong>and</strong> its downdip equivalents (Figs. 4, 6, 7). It onlaps<br />

much of the ramp, pinching out to a feather edge updip. The<br />

Hillsdale composite sequence thickens into the foredeep to 270 m<br />

(Bartlett <strong>and</strong> Webb, 1971; Al-Tawil, 1998), where it may consist of<br />

sequences dominated by shallow-water, fine-grained <strong>carbonate</strong>s<br />

<strong>and</strong> interbedded s<strong>and</strong>stone (not shown on cross sections in Figs.<br />

6, 7).<br />

Composite Sequence 2.—This composite sequence includes sequences<br />

C1 to 4, which make up the Denmar Limestone. It<br />

extends from the base of the basal red beds (<strong>and</strong> their downdip<br />

equivalents) of sequence C1, up to the base of the lower Taggard<br />

red beds (<strong>and</strong> downdip equivalents) of sequence C5 (Figs. 4, 6, 7).<br />

The lowst<strong>and</strong> systems tract is made up of the lowst<strong>and</strong> siliciclastics<br />

of sequence C1. The transgressive systems tract probably includes<br />

the rest of sequences C1 to C3. The maximum flooding<br />

surface is likely at the base of sequence C4 updip, <strong>and</strong> beneath the<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

C4 deeper-water mudstones downdip. The highst<strong>and</strong> systems<br />

tract is made up of sequence C4.<br />

25<br />

Composite Sequence 3.—Sequences C5 to 10 may be the third<br />

composite sequence, extending from the basal C5 red beds up to<br />

the top of C10 (Figs. 4, 6, 7). It is composed of the Pickaway, Union,<br />

<strong>and</strong> Alderson Limestones. The lowst<strong>and</strong> systems tract of the<br />

composite sequence consists of sequences C5 <strong>and</strong> C6, which are<br />

red-bed prone, <strong>and</strong> show a major basinward shift in both distribution<br />

<strong>and</strong> facies. The transgressive systems tract sequence of the<br />

composite consists of sequence C7 <strong>and</strong> the lower part of sequence<br />

C8. The maximum flooding surface is marked by the most l<strong>and</strong>ward<br />

extent of open-marine skeletal wackestone in sequence C8.<br />

There is a significant development of siliciclastics at the base of<br />

C9, including an updip tongue of red beds (Fig. 6, sections 1, 2).<br />

The highst<strong>and</strong> systems tract is made up of the upper half of<br />

sequences C8, C9, <strong>and</strong> C10.<br />

Composite Sequence 4.—This composite sequence likely extends<br />

from the base of the incised Fido S<strong>and</strong>stone–Lillydale Shale <strong>and</strong><br />

includes sequence C11 (Lillydale–Glenray interval); the top of the<br />

composite sequence is above the study interval (Figs. 4, 6, 7). The<br />

lowst<strong>and</strong> systems tract may include the incised siliciclastic units at<br />

the base of C11, <strong>and</strong> the maximum flooding surface is above the<br />

thin limestones at the base of C11 <strong>and</strong> below the fossiliferous<br />

shales. The highst<strong>and</strong> systems tract probably includes the Lillydale<br />

Shale, the Glenray Limestone, the Bickford Shale, <strong>and</strong> the Reynolds<br />

Limestone, <strong>and</strong> possibly units above the study interval.<br />

CONTROLS ON SEQUENCE DEVELOPMENT<br />

ON THE FORELAND<br />

Tectonics <strong>and</strong> Subsidence Rates<br />

Accommodation rates calculated from the present-day stratigraphic<br />

thicknesses (not decompacted) were highest (at least 10<br />

to 30 cm/ky) in the basin compared to the ramp margin (a<br />

minimum of 5 to 15 cm/ky) <strong>and</strong> were lowest (at least 1 to 3 cm/<br />

ky) in updip areas of the distal forel<strong>and</strong>. This promoted regional<br />

slopes into the forel<strong>and</strong> basin. Regional slopes per sequence can<br />

be calculated using the difference in thickness of the study<br />

interval from updip to downdip, <strong>and</strong> dividing this by the distance<br />

times the number of sequences. This gives slopes of at least<br />

5 cm/km/sequence on the distal forel<strong>and</strong> (Al-Tawil <strong>and</strong> Read,<br />

this volume) <strong>and</strong> average slopes of at least 20 cm/km/sequence<br />

down the plunging basin axis. The low distal forel<strong>and</strong> slopes<br />

resulted in the predominance of shallow-water facies over the<br />

whole of the distal forel<strong>and</strong>. Calculated slopes per sequence<br />

measured over the palinspastically restored dip direction (northwest<br />

to southeast) down the ramp margin <strong>and</strong> into the basin were<br />

higher, at least 30 cm/km. These calculated slopes are compatible<br />

with minimum estimated highst<strong>and</strong> water depths of at least 50 m<br />

on the ramp slope, which periodically shallowed to sea level<br />

during lowst<strong>and</strong>s.<br />

It has been suggested that the region was tectonically quiescent<br />

during deposition of the study interval, <strong>and</strong> in a state of<br />

elevational equilibrium between the filled forel<strong>and</strong> basin <strong>and</strong><br />

the beveled Devonian orogenic highl<strong>and</strong>s (Ettensohn, 1994).<br />

However, this appears to be incompatible with the considerable<br />

differential subsidence shown by the thickening of the <strong>carbonate</strong>s<br />

from a feather edge updip on the forel<strong>and</strong> to over 900 m in<br />

the basin (Figs. 6, 7). The thickening takes place not only across<br />

the regional hinge but also at various localities updip <strong>and</strong><br />

downdip from the hinge at various times. This complex pattern<br />

may have resulted from the movement of basement fault blocks


26<br />

beneath the forel<strong>and</strong>, which parallel the orogenic front<br />

(Shumaker <strong>and</strong> Wilson, 1996). However, given the long wavelengths<br />

of deformation, it also could just relate to flexural<br />

downwarping. Such downwarping in the forel<strong>and</strong> basin over<br />

the Price delta system occurred too long (some 10 to 15 My) after<br />

the initial early Mississippian loading event (Ettensohn, 1994) to<br />

be related to relaxation. This suggests that accommodation<br />

space might have been initiated during the Meramecian <strong>and</strong><br />

early Chesterian by renewed thrust loading followed by relaxation<br />

(cf. Quinlan <strong>and</strong> Beaumont, 1984; Al-Tawil <strong>and</strong> Read, this<br />

volume). This tectonic loading <strong>and</strong> differential subsidence during<br />

deposition of the Meramecian sequences <strong>and</strong> Chesterian<br />

sequences C1 to C6 caused rapid thickness changes (Ettensohn,<br />

1975, 1980; MacQuown <strong>and</strong> Pear, 1983; Dever, 1986, Dever et al.,<br />

1990; Shumaker <strong>and</strong> Wilson, 1996). In addition, there was a<br />

northwestward migration (approximately 50 km) of the major<br />

hinge onto the distal forel<strong>and</strong> as well as a northeastward migration<br />

of smaller hinge zones during the Meramecian–Chesterian<br />

(Figs. 6, 7). This would seem to support continued advance of a<br />

thrust (or sediment) load onto the forel<strong>and</strong>, suggesting that<br />

Greenbrier deposition occurred during active tectonics rather<br />

than during a time of tectonic quiescence.<br />

The regional subsidence that initiated Chesterian deposition,<br />

coupled with long-term sea-level rise (Ettensohn, 1994), moved<br />

the siliciclastic shorelines far updip during fourth-order<br />

highst<strong>and</strong>s, resulting in widespread <strong>carbonate</strong> deposition on the<br />

ramp <strong>and</strong> siliciclastic deposition only during fourth-order<br />

lowst<strong>and</strong>s/early transgressions. The marked differential subsidence<br />

exhibited by the Meramecian sequence <strong>and</strong> Chesterian<br />

sequences C1 to C6 (Figs. 6, 7) contrasts with the more gradual<br />

thickness changes of the upper sequences C7 to C10, which<br />

subsided more as a result of regional flexuring, possibly as<br />

tectonic loading slowed <strong>and</strong>/or regional uplift by erosional unloading<br />

of the orogenic highl<strong>and</strong>s was initiated. Perhaps the<br />

earlier faults on the forel<strong>and</strong> had locked by this time <strong>and</strong> the<br />

forel<strong>and</strong> underwent regional flexure. Rapid flexure over the<br />

hinge also characterizes sequence C11 (Figs. 6, 7) <strong>and</strong> sequence<br />

C12 (above the study interval) <strong>and</strong> may mark renewed loading of<br />

the forel<strong>and</strong>. Following deposition of the Chesterian <strong>carbonate</strong>s,<br />

erosional unloading <strong>and</strong> uplift of the orogenic highl<strong>and</strong>s caused<br />

the subsequent progradation of Upper Mississippian Mauch<br />

Chunk siliciclastic units (Ettensohn, 1994).<br />

Differential subsidence thus controlled thickness <strong>and</strong> facies<br />

variations across the ramp, not only across the major ramp<br />

margin hinge but also at various places on the distal forel<strong>and</strong>, <strong>and</strong><br />

thus was an important control on the accommodation created<br />

during deposition of each sequence. Broad sags periodically<br />

developed on the forel<strong>and</strong> in the northeast (sequences C2, C6, C7;<br />

Fig. 6) <strong>and</strong> locally on the distal forel<strong>and</strong> in the west (as in<br />

Meramecian sequence 2, <strong>and</strong> sequences C2, C3, <strong>and</strong> C9; Fig. 7).<br />

These sags <strong>and</strong> adjacent highs may have influenced the position<br />

of updip shoal-water grainstones separated by lagoonal lime<br />

wackestone–mudstone from the ramp-margin grainstones along<br />

the hinge. Although the differential subsidence modified the<br />

magnitudes of the relative sea-level changes, it did not override<br />

the eustatic changes that formed the sequence-bounding,<br />

siliciclastic-prone units <strong>and</strong> the <strong>carbonate</strong>-prone transgressive<br />

<strong>and</strong> highst<strong>and</strong> units.<br />

Climate<br />

Climate Control on Facies.—During the early Mississippian,<br />

the Appalachian Basin had a humid climate, which favored<br />

siliciclastic deposition on the Price–Borden delta complex<br />

(Branson, 1912; Butts, 1940, 1941; Sable <strong>and</strong> Dever, 1990). This<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

climate subsequently became more arid during deposition of<br />

the Maccrady red beds <strong>and</strong> evaporites. Evaporites in the<br />

Meramecian interval in Kentucky (southwest of the study area)<br />

indicate that arid <strong>and</strong> semiarid conditions persisted throughout<br />

the middle Mississippian (Sable <strong>and</strong> Dever, 1990; Khetani <strong>and</strong><br />

Read, 2002). This semiarid climate persisted into the early<br />

Chesterian during deposition of the lower part of the study<br />

interval (indicated by red beds, widespread oolitic ramps, coastal<br />

eolianites, <strong>and</strong> caliche in West Virginia, Kentucky, <strong>and</strong> the<br />

Illinois Basin (Dever, 1973, 1995; Ettensohn et al., 1988; Smith<br />

<strong>and</strong> Read, 1999). However, the climate may have become more<br />

humid into the later Chesterian, when disconformities in the<br />

Appalachian <strong>and</strong> Illinois basins became characterized by<br />

paleokarst (Ettensohn et al., 1988), coarse siliciclastic influx<br />

increased (Swann, 1964; Smith, 1996), skeletal limestones became<br />

more widespread, <strong>and</strong> thick oolitic deposits (which on<br />

ramps require semi arid conditions; Read, 1985) became much<br />

less common. Also, regional meteoric aquifers developed during<br />

this late Mississippian–Early Pennsylvanian humid phase<br />

(Niemann <strong>and</strong> Read, 1988). These climatic changes are borne<br />

out by paleogeographic reconstructions (de Witt <strong>and</strong> McGrew,<br />

1979; Scotese <strong>and</strong> McKerrow, 1990) that show the North American<br />

plate gradually rotating counterclockwise <strong>and</strong> moving northward<br />

from the Pangean arid sub-equatorial belt during the<br />

Meramecian to Chesterian (latitudes of 25 degrees or less),<br />

reaching close to the Equator by Pennsylvanian time.<br />

By latest Mississippian time, semiarid climate returned at<br />

least intermittently with the deposition of Mauch Chunk red beds<br />

<strong>and</strong> semiarid paleosols (Caudill et al., 1996).<br />

Climate Control on Long-Term Global Sea Levels.—Oxygen<br />

stable-isotope data from the mid-continent United States show<br />

several major climate events in the Meramecian–Chesterian,<br />

with increased δ 18 O values being interpreted as cooling <strong>and</strong>/or<br />

ice buildup events <strong>and</strong> decreased δ 18 O values representing<br />

warming events (Mii et al., 1999). There is a mid-Meramecian<br />

cooling event that likely marks the regressive event at the base<br />

of the Upper Meramecian sequences. This was followed by<br />

warming in the early Chesterian, which promoted transgression<br />

<strong>and</strong> deposition of the Meramecian sequences <strong>and</strong> Chesterian<br />

sequences C1 to C4. This was followed by a cooling event that<br />

likely corresponds to deposition of the red beds of sequences C5<br />

<strong>and</strong> C6 (Taggard red beds). Subsequent warming probably<br />

caused the long-term transgression <strong>and</strong> deposition of sequences<br />

C7 through C10. The subsequent cooling trend (Mii et al., 1999;<br />

Veevers <strong>and</strong> Powell, 1987) <strong>and</strong> falling sea levels were accompanied<br />

by progradation of Mauch Chunk siliciclastic-prone sequences<br />

(C11 <strong>and</strong> younger units).<br />

Eustasy<br />

Evidence for Fourth-Order Eustasy.—A set of relative-sea-level<br />

curves for the basin (Fig. 8) were constructed by arbitrarily<br />

assigning equal time to each of the sequences <strong>and</strong> plotting the<br />

relative magnitude of updip–downdip migration of marker facies<br />

such as oolitic or skeletal grainstones <strong>and</strong> s<strong>and</strong>stones. The<br />

resultant curves were compared to the global sea-level curves of<br />

Ross <strong>and</strong> Ross (1987) <strong>and</strong> a simplified Illinois Basin curve modified<br />

from Smith (1996) <strong>and</strong> Smith <strong>and</strong> Read (2000) from which the<br />

higher-frequency fluctuations were removed. In addition to the<br />

biostratigraphic control based on conodonts <strong>and</strong> crinoids<br />

(Collinson et al., 1971; Huggins, 1983; deWitt <strong>and</strong> McGrew, 1979;<br />

Sable <strong>and</strong> Dever, 1990), correlation between the Appalachian <strong>and</strong><br />

the Illinois basins was aided by the presence of relatively regional<br />

unconformities <strong>and</strong> siliciclastic markers in Kentucky (McFarlan


VISEAN NAMURIAN<br />

CHESTERIAN<br />

MERAMECIAN<br />

ST. LOUIS<br />

<strong>and</strong> Walker, 1956) <strong>and</strong> in West Virginia; such markers include the<br />

Taggard red beds, <strong>and</strong> the Greenville <strong>and</strong> Lillydale shales (Smith<br />

<strong>and</strong> Read, 1999, 2000, 2001).<br />

The Upper Meramecian sequences in the proximal forel<strong>and</strong><br />

correlate with the upper <strong>and</strong> lower St. Louis Limestone of the<br />

Illinois Basin (Huggins, 1983) but only one of the Upper<br />

Meramecian Hillsdale sequences appears to reach the distal<br />

forel<strong>and</strong> in Kentucky (Al-Tawil <strong>and</strong> Read, this volume).<br />

Most of the Chesterian fourth-order sequences in eastern<br />

West Virginia can be traced into much thinner disconformitybounded<br />

sequences exposed on the distal forel<strong>and</strong>, eastern Kentucky<br />

(Al-Tawil, 1998; Al-Tawil <strong>and</strong> Read, this volume), <strong>and</strong> into<br />

the Illinois Basin (Ettensohn et al., 1984; Smith, 1996; Smith <strong>and</strong><br />

Read, 1999; 2001) (Fig. 8).<br />

The four West Virginia–Virginia sequences (C1 to C4) in the<br />

Playtycrinites penicillus–bearing Denmar interval (G. bilineatus–C.<br />

charactus conodont zone; Huggins, 1983; Stamm, 1997) can be<br />

recognized in the time-equivalent Ste. Genevieve Limestone,<br />

exposed on the distal forel<strong>and</strong> in Kentucky. However, they are<br />

difficult to match with the three Ste. Genevieve sequences in the<br />

Illinois Basin, although four sequences may occur there if the two<br />

disconformity-bounded units in Smith <strong>and</strong> Read’s (1999) sequence<br />

C3 are raised to sequence status.<br />

West Virginia sequence C5 (containing the lower Taggard<br />

red beds), sequences C6 (bearing the upper Taggard red beds–<br />

lower Pickaway) <strong>and</strong> C7 of West Virginia (upper Pickaway<br />

interval), which contain the first occurrences of the crinoid<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

ST. LOUIS<br />

STE. GENEVIEVE<br />

APPALACHIAN BASIN<br />

GLOBAL SEA LEVEL<br />

ROSS AND ROSS (1987)<br />

ILLINOIS BASIN<br />

SMITH (1996)<br />

KENTUCKY<br />

DISTAL FORELAND<br />

(AL-TAWIL AND READ,<br />

THIS VOLUME)<br />

WEST VIRGINIA<br />

PROXIMAL FORELAND<br />

(THIS STUDY)<br />

HIGH LOW HIGH LOW HIGH LOW HIGH LOW<br />

GLEN DEAN<br />

GLEN DEAN 10<br />

GLEN DEAN<br />

HARDINSBURG SS.<br />

C11<br />

C11 GLEN RAY<br />

HANEY<br />

HANEY 9<br />

HANEY C10<br />

C10 ALDERSON<br />

BEECH CK<br />

BEECH CK 8<br />

CYPRESS SS<br />

BEECH CK C9<br />

CAVE BRANCH SH.<br />

C9 UPPER UNION<br />

REELSVILLE 7<br />

SAMPLE SS.<br />

BEAVER BEND 6<br />

BETHEL SS.<br />

REELSV. C8B<br />

C8A<br />

BEAVER<br />

BEND<br />

C8B<br />

C8A<br />

LOWER<br />

UNION<br />

5<br />

C7<br />

C7<br />

PAOLI<br />

PAOLI<br />

C6<br />

C6<br />

4<br />

BRYANTSVILLE BRECCIA<br />

3 LEVIAS AUX<br />

VASE<br />

C5<br />

C4<br />

C5<br />

C4<br />

U. TAGGARD<br />

L. TAGGARD<br />

STE. GENEVIEVE<br />

2<br />

SPAR MT.<br />

SS<br />

C3<br />

STE. GENEVIEVE<br />

C3<br />

DENMAR<br />

1<br />

C2<br />

C2<br />

PAOLI /<br />

WARIX RUN<br />

C1<br />

ST. LOUIS<br />

HILLSDALE<br />

FIG. 8.—Relative third-order <strong>and</strong> fourth-order sea-level curves for the study interval in the West Virginia–Virginia Appalachians<br />

compared with curves from the distal Appalachian forel<strong>and</strong> in Kentucky (Al-Tawil <strong>and</strong> Read, this volume), the Illinois Basin<br />

(Smith, 1999; 2000; 2001), <strong>and</strong> the third-order curves of Ross <strong>and</strong> Ross (1987). Most of the fourth-order cycles can be correlated<br />

between the basins. The Ross <strong>and</strong> Ross third-order curves can be tied grossly to the Appalachian third-order cycles. The Illinois<br />

Basin curves show some additional large sea-level falls defined on the degree of incision, which are not recognized in the<br />

Appalachian Basin.<br />

27<br />

Talarocrinus, are correlated with the three sequences mapped as<br />

“Warix Run”/Paoli in Kentucky by Ettensohn et al. (1984), <strong>and</strong><br />

the Talarocrinus-bearing Paoli sequences in the Illinois Basin<br />

(Butts, 1922, 1940, 1941; Wells, 1950; Ettensohn et al., 1984; Smith<br />

<strong>and</strong> Read, 1999).<br />

West Virginia sequence C8 (Lower Union) corresponds to the<br />

Beaver Bend–Reelsville cycle in Kentucky <strong>and</strong> in the Illinois Basin<br />

(de Witt <strong>and</strong> McGrew, 1979; Ettensohn et al., 1984; Smith <strong>and</strong><br />

Read, 2001; Al-Tawil <strong>and</strong> Read, this volume). Sequence C9 of<br />

West Virginia (mid-Union red beds <strong>and</strong> Upper Union Limestone),<br />

containing G. bilineatus–C. altus conodonts, was correlated<br />

with the Reelsville limestone of the Illinois Basin (de Witt <strong>and</strong><br />

McGrew, 1979), but it more likely correlates with the Cave Branch<br />

Shale–Beech Creek sequence in Kentucky <strong>and</strong> the Cypress S<strong>and</strong>stone–Beech<br />

Creek–Big Clifty sequence (Illinois Basin), containing<br />

basal G. bilineatus–C. altus zonal conodonts (Collinson et al.,<br />

1971; Ettensohn et al., 1984).<br />

West Virginia sequence C10 (Greenville Shale–Alderson Limestone),<br />

containing the up<strong>permo</strong>st occurrence of Talarocrinus (Butts,<br />

1940; Wells, 1950), correlates with the Haney sequence of Kentucky<br />

<strong>and</strong> the Illinois Basin (de Witt <strong>and</strong> McGrew, 1979; Ettensohn<br />

et al., 1984; Smith <strong>and</strong> Read, 2001; Al-Tawil <strong>and</strong> Read, this<br />

volume). West Virginia sequence C11 (Lillydale Shale/Fido S<strong>and</strong>stone–Glenray<br />

Limestone), which contains the G. bilineatus–K.<br />

mehli zonal conodonts (Rexroad <strong>and</strong> Clark, 1960) correlates with<br />

the Hardinsburg–Glen Dean cycle of Kentucky <strong>and</strong> the Illinois<br />

Basin (de Witt <strong>and</strong> McGrew, 1979; Ettensohn et al., 1984).<br />

C1<br />

PICKAWAY


28<br />

The interbasinal extent of these sequences suggests that global<br />

eustasy played a major role in their development, rather than<br />

tectonics alone, which would operate locally as the basins subsided<br />

differentially. Although better age constraints are needed,<br />

the high-frequency sequences appear to relate to fourth-order<br />

sea-level cycles. If with better dating these do prove to be 400 ky,<br />

then they may relate to Milankovitch-driven, long-term eccentricity,<br />

which would be expected to predominate during times of<br />

global ice sheets (Imbrie <strong>and</strong> Imbrie, 1986; Ruddiman et al., 1986).<br />

This is compatible with the onset of late Paleozoic Gondwana<br />

glaciation in the Southern Hemisphere, in the Early Chesterian or<br />

even earlier (Horbury, 1989; Frakes et al., 1992; Smith <strong>and</strong> Dorobek,<br />

1993; Wright <strong>and</strong> Vanstone, 2001).<br />

The data from the Illinois <strong>and</strong> Appalachian basins suggest<br />

that there was an increase in the magnitude of the sea-level<br />

changes through the Chesterian, peaking in the Pennsylvanian<br />

with sea-level changes of up to 100 m (Heckel, 1986). However,<br />

the presence of significant incision on some of the basal Chesterian<br />

units in the Illinois Basin may suggest that there were some large<br />

sea-level falls even this early (Leetaru, 2000; Smith <strong>and</strong> Read,<br />

2000, 2001) <strong>and</strong> possibly earlier in the Meramecian (Smith <strong>and</strong><br />

Dorobek, 1993; Wright <strong>and</strong> Vanstone, 2001; Horbury, 1989) (Fig.<br />

8). This upward increase in magnitude of Late Mississippian<br />

fourth-order sea-level cycles was superimposed on the long-term<br />

Chesterian transgression, resulting in progressive l<strong>and</strong>ward shift<br />

of fourth-order highst<strong>and</strong>s. However, the increasing sea-level<br />

falls kept returning the lowst<strong>and</strong> s<strong>and</strong> bodies to the general<br />

vicinity of the ramp margin <strong>and</strong> slope, even though long-term<br />

transgression was occurring.<br />

Evidence for Third-Order Eustasy.—Some of the weak, composite<br />

(third-order) sequences appear to correspond to third-order<br />

sea-level cycles documented as global events by Ross <strong>and</strong> Ross<br />

(1987) (Fig. 8). The third-order cycle that formed the Little Valley–<br />

Hillsdale composite sequence correlates with the Late Meramecian<br />

(St. Louis) sea-level cycle (Huggins, 1983). The third-order cycle<br />

that formed the Denmar composite sequence correlates with the<br />

Ste. Genevieve (earliest Chesterian) sea-level cycle, with a major<br />

lowst<strong>and</strong> at the position of the Appalachian Basin sequences C5<br />

<strong>and</strong> C6 red beds (lower <strong>and</strong> upper Taggard red beds). However,<br />

the later composite sequences are difficult to relate to the global<br />

cycles, in part because of insufficient detail in the Ross <strong>and</strong> Ross<br />

(1987) cycle chart. However, they do show a third-order sea-level<br />

cycle extending from the top of the Ste. Genevieve cycle up to the<br />

top of the Glen Dean (top of our C11), whereas we placed the top<br />

of the cycle at the base of C11 (base of Glenray).<br />

There also are substantial differences between the major<br />

incisional events in the Appalachian <strong>and</strong> Illinois basins. In the<br />

Illinois Basin, Leetaru (2000) recognized major incision in the<br />

middle <strong>and</strong> near the top of the Ste. Genevieve <strong>carbonate</strong>s, whereas<br />

we see major lowst<strong>and</strong>s only above the equivalent (Denmar)<br />

<strong>carbonate</strong>s in the Appalachian Basin (at bases of C5 <strong>and</strong> C6).<br />

Incisional events in the Illinois Basin (Smith <strong>and</strong> Read, 2000, 2001)<br />

for which there is little evidence in the Appalachian Basin also<br />

include the Bethel S<strong>and</strong>stone incision (base of our C8 sequence),<br />

<strong>and</strong> the incision at the base of the Cypress S<strong>and</strong>stone, Illinois<br />

Basin (base of our C9), although there is a red-bed tongue at this<br />

level updip in the Appalachian Basin. Incision at the base of the<br />

Hardinsburg S<strong>and</strong>stone, in the Illinois Basin, corresponds to the<br />

incision observed at the base of our sequence C11 (Fido S<strong>and</strong>stone).<br />

The differences between the amount of incision between<br />

the Appalachian <strong>and</strong> the Illinois basins (Smith <strong>and</strong> Read, 1999,<br />

2000, 2001) may reflect the effects of uplift or subsidence acting on<br />

weak third-order sea-level changes, coupled with siliciclastic<br />

influx <strong>and</strong> incision associated with significant fluvial systems in<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

the Illinois Basin, which may have accentuated the effects of some<br />

of the eustatic falls in the Illinois Basin. The greater subsidence<br />

rates at least in the proximal forel<strong>and</strong> of the Appalachian Basin<br />

may have tended to inhibit incision, as did the relative paucity of<br />

siliciclastic systems there.<br />

Development of Fourth-Order Sequences<br />

Development of Sequence Boundaries.—The disconformities <strong>and</strong><br />

associated caliches <strong>and</strong> breccias that mark sequence boundaries<br />

updip in Kentucky as well as in the Illinois Basin (Ettensohn et al.,<br />

1984; Ettensohn et al., 1988; Smith <strong>and</strong> Read, 1999, 2001; Al-Tawil<br />

<strong>and</strong> Read, this volume), appear to be less developed in the West<br />

Virginia study area, although this may be more apparent than<br />

real <strong>and</strong> may reflect the dominance of well data over outcrop data<br />

in this paper. However, it could also be due to the rapid mantling<br />

of exposed <strong>carbonate</strong>s by siliciclastic veneers (eolian or sheetwash<br />

deposition) that inhibited paleosol formation. Local,<br />

calichified <strong>and</strong> brecciated units downdip along the ramp margin<br />

may reflect the low gradient of the ramp top, which allowed it to<br />

be exposed during sea-level falls. Sharp, planar to hummocky<br />

erosion surfaces that are developed on some of the grainstones at<br />

parasequence <strong>and</strong> sequence boundaries in outcrops downdip<br />

could reflect widespread marine cementation followed by erosion<br />

in rock platform settings during transgression. Significant<br />

meteoric cementation beneath disconformities did not occur<br />

until later in the Mississippian, when the climate became more<br />

humid (Niemann <strong>and</strong> Read, 1988).<br />

Development of Lowst<strong>and</strong> to Early Transgressive Tracts.—Although<br />

regional 3-D data are needed, it appears that siliciclastics<br />

for the lowst<strong>and</strong>/early transgressive s<strong>and</strong> bodies were sourced<br />

from highl<strong>and</strong>s to the northeast <strong>and</strong> cratonic sources to the north,<br />

<strong>and</strong> they traveled southward toward the ramp margin (Fig. 1). It<br />

is not yet clear whether any were derived from the highl<strong>and</strong>s<br />

bordering the eastern side of the foredeep directly east of the<br />

study area, inasmuch as this either would have required very<br />

large sea-level falls to allow transport across the foredeep. The<br />

distribution of regional red beds in sequences C5 <strong>and</strong> C6 (Taggard<br />

red beds) suggests that siliciclastics from the northeast were<br />

funneled down the eastern side of the forel<strong>and</strong>, which may have<br />

been a subtle topographic low that deepened into the foredeep to<br />

the southwest. Sequences C5 <strong>and</strong> C6 red beds were not deposited<br />

on the topographically higher (?) distal forel<strong>and</strong> to the west,<br />

which was the site of eolian <strong>and</strong> early transgressive shallow<br />

marine siliciclastic deposition. Siliciclastic veneers at bases of the<br />

sequences on the distal forel<strong>and</strong> could have been transported<br />

onto the ramp during lowst<strong>and</strong>, by winds or flash floods, but<br />

most of the siliciclastic units probably were subsequently reworked<br />

into early transgressive marine sheetlike units. Quartz–<br />

peloid eolianites may have been preserved (as in the outcropping<br />

distal forel<strong>and</strong> deposits of Kentucky; Al-Tawil <strong>and</strong> Read, this<br />

volume), perhaps because of early incipient cementation, which<br />

inhibited reworking, but others could have been reworked into<br />

marine quartz–peloid grainstone sheets, which would be difficult<br />

to differentiate from eolian deposits in cuttings. Some of the<br />

lowst<strong>and</strong> to early transgressive s<strong>and</strong>stone–calcareous siltstones<br />

were deposited as barriers behind which thin, nearshore shales<br />

were deposited, as in sequences C7 to C10.<br />

Because of the poor constraints on the water depths of the<br />

facies, estimates of sea-level fall forming the lowst<strong>and</strong> deposits are<br />

crude. Where relative sea-level falls were small to moderate (say 10<br />

to 15 m or so), lowst<strong>and</strong> marine s<strong>and</strong>stones that are only slightly<br />

shallower than the underlying highst<strong>and</strong> facies were deposited,<br />

<strong>and</strong> there was only a small basinward shift in facies. These include


lowst<strong>and</strong>s at the bases of sequences C2, C7, <strong>and</strong> C8. Where relative<br />

sea-level falls were larger (say 20 to 40 m or so), then forced<br />

regression caused a significant basinward shift of facies, <strong>and</strong><br />

lowst<strong>and</strong> shallow-water ooid grainstones or marine s<strong>and</strong>stones<br />

were deposited on deeper-water skeletal wackestone or laminated<br />

shaly lime mudstone, with intervening facies not being deposited,<br />

as at the bases of sequences C1, C6, C9, C10, <strong>and</strong> C11.<br />

The localization of lowst<strong>and</strong> wedges on the ramp margin was<br />

controlled by high differential subsidence over the hinge <strong>and</strong> onto<br />

the proximal forel<strong>and</strong> during deposition of the lower sequences.<br />

During deposition of sequences C1 to C6 (Denmar <strong>and</strong> lower<br />

Pickaway sequences), where the underlying highst<strong>and</strong> skeletal<br />

bank facies were located close to the ramp margin, lowst<strong>and</strong> s<strong>and</strong>s<br />

migrated onto the ramp slope as sea level fell. However, during<br />

deposition of sequences C7 <strong>and</strong> C8, the ramp margin was farther<br />

updip, as a probable response to the long-term sea-level rise <strong>and</strong><br />

activation of hinges farther updip. Thus, sea level had to fall farther<br />

to expose the ramp margin <strong>and</strong> slope, so that lowst<strong>and</strong> siliciclastics<br />

reached only the ramp margin. However, the increasingly large<br />

sea-level falls during deposition of sequences C9 to C11, coupled<br />

with the broad flexuring of the ramp, rather than differential<br />

subsidence, could have returned deposition of lowst<strong>and</strong> quartz<br />

s<strong>and</strong>stone <strong>and</strong> oolite units to the ramp slope. The most well<br />

developed lowst<strong>and</strong> siliciclastic units appear to coincide with<br />

third-order sea-level falls, as in the base of sequence C1 (base of the<br />

Denmar), the bases of sequences C5 <strong>and</strong> C6 (Taggard red beds),<br />

<strong>and</strong> the base of sequence C11 (Fido S<strong>and</strong>stone).<br />

Transgressive Systems Tract Development <strong>and</strong> Maximum Flooding.—Although<br />

the maximum flooding surfaces of each of the<br />

sequences were difficult to pick at all locations, given the mix of<br />

subsurface well data <strong>and</strong> higher-quality outcrop data, some<br />

generalizations can be made. Along the ramp slope, the coincidence<br />

of maximum flooding surfaces with the transgressive<br />

surface on the lowst<strong>and</strong> systems tract (as in sequences C 4, C10<br />

<strong>and</strong> C11) appears to reflect rapid deepening <strong>and</strong> drowning of the<br />

lowst<strong>and</strong> facies as relative sea level rose <strong>and</strong> the ramp was<br />

transgressed. In contrast, in most of the other sequences, transgressive<br />

wackestone units, or transgressive–regressive skeletal<br />

wackestone to skeletal grainstone (C1, C3, C6) were developed on<br />

the lowst<strong>and</strong> deposits, reflecting more gradual relative-sea-level<br />

rise during fourth-order transgression. If the skeletal units on the<br />

ramp slope developed solely under the influence of gradually<br />

rising sea level, they would be expected to be upward deepening<br />

into the overlying deeper-water facies. However, the skeletal<br />

banks appear to be shallowing-upward successions <strong>and</strong> probably<br />

reflect superimposed higher-frequency sea-level fluctuations<br />

or perhaps a slowing of subsidence.<br />

On the shallow ramp, some fourth-order sea-level rises reworked<br />

basal siliciclastic units but did not locally deposit a<br />

transgressive succession of <strong>carbonate</strong>s, although such transgressive<br />

deposits formed within some of these same sequences elsewhere.<br />

Suppression of transgressive <strong>carbonate</strong> deposition could<br />

have been due to rapid rise rates that did not allow time for a<br />

significant transgressive <strong>carbonate</strong> to be deposited, or perhaps<br />

the shallow marine waters were too turbid with resuspended<br />

terrigenous silt. Deposition of a transgressive <strong>carbonate</strong> succession<br />

on the locally developed, eolian <strong>and</strong> marine reworked<br />

siliciclastic units could have been favored by slower rise of rates<br />

of relative sea level <strong>and</strong> waters becoming clearer as transgression<br />

proceeded. The presence of upward-shallowing units that make<br />

up the transgressive <strong>carbonate</strong>s in some sequences that have<br />

significant transgressive systems tracts (sequence C2, C7, <strong>and</strong> C8;<br />

Fig. 6), suggests that transgressive <strong>carbonate</strong> deposition may<br />

have been favored by a short-term rise <strong>and</strong> fall superimposed on<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

the fourth-order rise, which effectively delayed maximum flooding<br />

of the ramp.<br />

29<br />

Highst<strong>and</strong> Deposition.—During highst<strong>and</strong> deposition of the<br />

sequences, a complex facies mosaic developed, rather than a<br />

simple upward-shallowing, progradational succession that would<br />

result from seaward migration of shore-parallel facies belts.<br />

Skeletal grainstone–packstone <strong>and</strong> ooid grainstone were deposited<br />

not only along the ramp margin but also as extensive but<br />

discontinuous bodies far back from the margin, from which they<br />

were separated by extensive lagoonal (<strong>and</strong> locally tidal flat) lime<br />

wackestone–mudstone. Some of these updip skeletal <strong>and</strong> ooid<br />

bodies could have formed bordering highs adjacent to locally<br />

downwarped areas of the forel<strong>and</strong> (Carney, 1993), but others<br />

could have been localized by subtle depositional topography.<br />

Individual ramp-margin ooid bodies were shaped by tidal currents<br />

in a northeast–southwest direction normal to depositional<br />

strike (Kelleher <strong>and</strong> Smosna, 1993).<br />

Lagoonal lime wackestone–mudstones were widely deposited<br />

in protected lagoons behind ramp-margin ooid <strong>and</strong> skeletal<br />

grainstone shoals <strong>and</strong> far updip between <strong>and</strong> behind shoals. The<br />

absence of tidal-flat laminites from much of the distal forel<strong>and</strong><br />

probably was due to the relatively large sea-level falls, which<br />

caused rapid seaward shift of the shoreline, faster than tidal flats<br />

could prograde. However, locally thick tidal-flat facies were<br />

deposited just behind the ramp margin as the lagoonal muds<br />

shallowed upward, possibly because of high sedimentation rates<br />

here, coupled with the low rates of relative sea-level fall, which<br />

would characterize late highst<strong>and</strong>s, <strong>and</strong> the relatively high subsidence<br />

rates, which would have decreased the effects of eustatic<br />

fall.<br />

Development of Late Meramecian–Chesterian<br />

Composite Sequences<br />

In the study area, the Kinderhook–middle Osagean Borden–<br />

Price delta complex was not overlain by upper Osagean–lower<br />

Meramecian <strong>carbonate</strong>s of the Fort Payne–Salem supersequence,<br />

which were deposited farther to the south. Thus, the upper<br />

Meramecian–Chesterian supersequence in this paper was deposited<br />

directly onto the earlier Price–Borden siliciclastic units, following<br />

rapid differential subsidence of the delta in the east,<br />

coupled with global sea-level rise (Dennison, 1985; Ettensohn,<br />

1994), which led to increased accommodation <strong>and</strong> long-term<br />

onlap of the upper Meramecian <strong>and</strong> Chesterian sequences onto<br />

the delta complex.<br />

Differential subsidence over the eastern hinge, coupled with<br />

relatively small third-order sea-level fluctuations which started<br />

from the low sea-level position at the end of the upper Osagean–<br />

lower Meramecian supersequence, limited deposition of composite<br />

sequence C1 (Little Valley–Hillsdale) to the margin of the<br />

distal forel<strong>and</strong> to the west. Deposition extended farther updip in<br />

the east, where subsidence was more rapid. The small superimposed<br />

fourth-order sea-level changes periodically flooded the<br />

ramp to shallow depths, resulting in deposition of dominantly<br />

muddy, restricted <strong>carbonate</strong>s of composite sequence C1.<br />

The subsequent sea-level lowst<strong>and</strong> initiated localized red bed<br />

<strong>and</strong>/or eolianite deposition on the southern <strong>and</strong> eastern edge of<br />

the distal forel<strong>and</strong>, which continued to undergo differential<br />

subsidence over the hinge during the four transgressive–regressive<br />

events that formed composite sequence C2. Long-term relative<br />

sea-level rise caused progressive backstepping of the updip<br />

limits of sequences C1 to 4 onto the distal forel<strong>and</strong>, while considerable<br />

differential subsidence across the hinge promoted deepwater<br />

deposition in the southeast. The small, fourth-order sea-


30<br />

level changes only shallowly flooded the distal forel<strong>and</strong> to the<br />

west, with deposition of shallow lagoonal muds <strong>and</strong> intervening<br />

ooid shoals. More open marine skeletal facies were deposited in<br />

the more rapidly subsiding areas southeast of the hinge, <strong>and</strong> to<br />

the northeast updip along the basin axis.<br />

With long-term sea-level fall, widespread red beds <strong>and</strong><br />

eolianites were deposited, initiating deposition of composite<br />

sequence C3 (Taggard red beds). The earlier hinge became less<br />

active during C5 deposition but was reactivated during C6 deposition,<br />

along with differential subsidence seaward of a local hinge<br />

updip to the northeast. The sea-level rises from the initial lowst<strong>and</strong><br />

positions (base C5 <strong>and</strong> base C6) flooded the ramp only to shallow<br />

depths, so that restricted facies were deposited over much of the<br />

ramp. However, long-term sea-level rise, coupled with increasing-amplitude<br />

fourth-order eustasy, caused sequences C5 to C10<br />

to continue to onlap updip, with maximum flooding apparently<br />

occurring in the middle of sequence C8, followed by slight offlap<br />

during deposition of subsequent sequences (C9 <strong>and</strong> C10). These<br />

increased-amplitude fourth-order cycles, coupled with slight<br />

westward <strong>and</strong> northward migration of the hinges, promoted<br />

backstepping of ramp-margin facies <strong>and</strong> deposition of relatively<br />

open marine facies over the seaward edge of the distal forel<strong>and</strong><br />

during C7 to C8 <strong>and</strong> perhaps C9 deposition. The decrease in<br />

deposition of ooid grainstone from sequence C9 to C11 may be<br />

due to the increasingly humid climate. As amplitudes of sea-level<br />

change increased, increased sea-level falls allowed lowst<strong>and</strong><br />

siliciclastics <strong>and</strong> ooid grainstones to be deposited farther downdip<br />

during deposition of sequences C9 to base of C11.<br />

Composite sequence C4 was initiated with incision <strong>and</strong> deposition<br />

of s<strong>and</strong> bodies at the base of sequence C11, probably<br />

because of very low sea level. Increasingly humid conditions,<br />

coupled perhaps with uplift in the source areas to the northeast,<br />

generated the shale-prone sequences C11 <strong>and</strong> C12 (above the<br />

study interval). Thickening of at least sequence C11 into the<br />

foredeep may reflect a return to higher differential subsidence<br />

associated with the major hinge, culminating in thick, deep-water<br />

<strong>carbonate</strong>s along the basin margin. Carbonate for these must<br />

have been shed from thick ramp-margin bank complexes along<br />

the ramp margin in the southeast (exposed along Pine Mountain,<br />

Kentucky; Al-Tawil <strong>and</strong> Read, this volume). Renewed, relatively<br />

deep submergence during deposition of sequence C11 (Figs. 6, 7)<br />

probably reflects this increased subsidence coupled with the<br />

relatively high-amplitude fourth-order sea-level changes. Shaleprone<br />

sequence C11, as well as the overlying, <strong>carbonate</strong>-poor<br />

Mauch Chunk siliciclastic sequences above the study interval,<br />

formed as long-term sea level started to fall, global climate cooled<br />

(Mii et al., 1999), climate became more humid, <strong>and</strong> erosional<br />

unloading of the orogenic belt occurred (Ettensohn, 1994), all of<br />

which promoted terrigenous influx onto the ramp <strong>and</strong> siliciclastic<br />

progradation.<br />

Moderate-Amplitude Glacio-Eustasy<br />

<strong>and</strong> Late Mississippian Ramps<br />

The study interval spans the onset of the Southern Hemisphere<br />

Gondwana glaciation (Fig. 9), <strong>and</strong> the estimated<br />

periodicities of the sequences <strong>and</strong> the facies developed suggest<br />

that moderate-amplitude glacio-eustasy forced by long-term eccentricity<br />

(Horbury, 1989; Frakes et al., 1992; Smith <strong>and</strong> Dorobek,<br />

1993; Miller <strong>and</strong> Eriksson, 1997; Smith <strong>and</strong> Read, 2000, 2001;<br />

Wright <strong>and</strong> Vanstone, 2001) may have controlled the formation of<br />

the sequences. This moderate-amplitude glacio-eustasy, which<br />

increased in magnitude with time, was transitional into the highamplitude<br />

glacio-eustasy of the Pennsylvanian (Fischer, 1982;<br />

Goldhammer et al., 1991; Heckel, 1986). The moderate-amplitude<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

sea-level falls of the Late Mississippian caused large-scale basinward<br />

migration of siliciclastic shorelines, <strong>and</strong> caused lowst<strong>and</strong> to<br />

early transgressive siliciclastics to be deposited on the ramp,<br />

ramp margin, <strong>and</strong> ramp slope. The increasingly large, moderateamplitude<br />

sea-level rises superimposed on a long-term eustatic<br />

rise, <strong>and</strong> the load-induced forel<strong>and</strong> subsidence, caused extensive,<br />

periodic floodings of the ramp. However, the sea-level rises<br />

were not sufficiently large to drown the ramps, so that widespread<br />

grainstones were deposited not only on the ramp margin<br />

but also over large areas far updip on the ramp, where they<br />

interfinger laterally with muddy lagoonal <strong>carbonate</strong>s. Ramps<br />

that formed under such moderate-amplitude eustasy contrast<br />

with greenhouse ramps (Fischer, 1982), on which significant<br />

grainstone units tend to be limited to the outer part of the shallow<br />

ramp, <strong>and</strong> parasequences are meter-scale cycles dominated by<br />

restricted, muddy shallowing-upward facies (Read, 1995;<br />

Koerschner <strong>and</strong> Read, 1989; Goldhammer et al., 1990). At the<br />

other end of the spectrum, the bulk of the sequences of the<br />

Mississippian ramp in West Virginia lack the widespread, onramp,<br />

deeper-water maximum-flooding black shale–lime mudstone<br />

facies of the Pennsylvanian cyclothems, typical of higheramplitude<br />

glacio-eustasy during icehouse times (Heckel, 1986;<br />

Smith <strong>and</strong> Read, 2000, 2001).<br />

CONCLUSIONS<br />

This study illustrates how well cuttings <strong>and</strong> wireline logs,<br />

integrated with limited core <strong>and</strong> outcrop data, can be used to<br />

generate high-resolution sequence stratigraphies of Mississippian<br />

ramps in the subsurface. The Meramecian to Chesterian<br />

Greenbrier succession in the West Virginia portion of the Appalachian<br />

Basin formed on a tropical, mixed <strong>carbonate</strong>–siliciclastic<br />

ramp that developed on a subsiding forel<strong>and</strong> that was undergoing<br />

differential subsidence, during the initiation of significant<br />

Southern Hemisphere, Gondwana, glaciation. The sequence<br />

stratigraphy was analyzed using two regional cross sections, one<br />

a dip section <strong>and</strong> the other an axial section, which were based on<br />

subsurface well cuttings <strong>and</strong> wireline logs from numerous shallow<br />

wells, several detailed outcrop sections, <strong>and</strong> a core. The<br />

Greenbrier succession forms the transgressive part of the upper<br />

Meramecian–Chesterian supersequence, <strong>and</strong> is made up of fourthorder<br />

depositional sequences stacked into variably developed<br />

third-order, composite sequences. The fourth-order depositional<br />

sequences reach thicknesses of tens of meters to a hundred<br />

meters. They are bounded by lowst<strong>and</strong> to early transgressive<br />

siliciclastic units <strong>and</strong> local updip eolianites. Transgressive banks<br />

developed along the ramp margin locally, but on the shallow<br />

ramp, transgressive shallow-water <strong>carbonate</strong>s were recognized<br />

in some areas but not in others. Highst<strong>and</strong> facies are extensive<br />

lagoonal lime wackestone–mudstone, along with ooid grainstone<br />

<strong>and</strong> skeletal grainstone–packstone that developed along the ramp<br />

margin <strong>and</strong> as extensive isolated units on the ramp interior, <strong>and</strong><br />

constitute the bulk of many sequences. Skeletal wackestone grading<br />

seaward into laminated, argillaceous lime mudstone developed<br />

on the deeper ramp <strong>and</strong> ramp slope.<br />

Deposition of the succession was influenced by differential<br />

subsidence of the foredeep, possibly marking a phase of late<br />

Paleozoic thrust loading. Paleoslopes on the distal forel<strong>and</strong> were<br />

a few centimeters per kilometer, increasing to 30 cm/km over the<br />

hinge. Accommodation rates were a few centimeters per kiloyear<br />

on the distal forel<strong>and</strong> increasing to 10 to 30 cm/ky on the<br />

proximal forel<strong>and</strong>. Tectonic subsidence generated a basinwardthickening<br />

wedge of sediments 0 to 900 m thick. Rapid thickening<br />

of many of the sequences occurs across the basin hinge, as well as<br />

across broad downwarps on the ramp, which were intermittently


active. This differential subsidence generated a complex mosaic<br />

of facies on the ramp; thus, sequences rarely show simple upward-deepening<br />

to upward-shallowing successions, or uniformly<br />

basinward-thickening trends.<br />

The fourth-order sea-level fluctuations that formed the sequences<br />

likely relate to glacially driven Milankovitch long-term<br />

eccentricity forcing. They were responsible for development of<br />

the characteristic sequence stacking with basal siliciclastic-prone<br />

units <strong>and</strong> upper <strong>carbonate</strong>-prone units, because they caused<br />

large shifts in the shoreline across the ramp. Sea-level changes<br />

became larger through time, which, along with the increasingly<br />

humid climate, caused an upsection change in the supersequence<br />

from dominantly restricted, ooid grainstone–prone sequences to<br />

more open marine, skeletal limestone–prone sequences with<br />

significant siliciclastic units. Third-order sea-level falls appear to<br />

have accentuated fourth-order falls, causing widespread deposition<br />

of red beds <strong>and</strong> marine siliciclastic units, <strong>and</strong> some incision<br />

of underlying units. Although subsidence rates on the proximal<br />

forel<strong>and</strong> were an order of magnitude greater than those acting on<br />

the distal forel<strong>and</strong>, fourth-order sea-level fluctuations (indicated<br />

by interbasinal correlation) still were the prime cause of the<br />

sequences, whereas tectonics strongly controlled regional <strong>and</strong><br />

local thickness <strong>and</strong> facies variations.<br />

ACKNOWLEDGMENTS<br />

We would like to thank the West Virginia Economic <strong>and</strong><br />

Geological Survey <strong>and</strong> especially Lee Avery, Dave Matchen, Pat<br />

Johns, <strong>and</strong> staff, for cuttings, cores, maps, <strong>and</strong> digital data files. We<br />

give special thanks to Taury Smith, whose outst<strong>and</strong>ing Illinois<br />

Basin work helped clarify many of the problems of Mississippian<br />

facies <strong>and</strong> stratigraphy, as did the work of Garl<strong>and</strong> Dever of the<br />

Kentucky Geological Survey. Bill Henika (Virginia Division of<br />

Mineral Resources) provided valuable insight into problems of<br />

Appalachian geology. Jim Molloy, Art Hall, Sarah Crews, Steve<br />

Turpin, Jane Roa, John Leone, Chad Haiar, Susan Barbor, Pat<br />

Barnhart, Don Thorsteson, Ashley Goodrich, Julia Caldwell, <strong>and</strong><br />

Jen Hatfield helped collect <strong>and</strong> process data. Mike Pope, Anna<br />

Balog, Amy Khetani, Dan Miller, <strong>and</strong> Brian Coffey provided<br />

valuable discussion <strong>and</strong> help, <strong>and</strong> Ellen Mathena did much of the<br />

editing <strong>and</strong> typing. The study was supported by Petroleum Research<br />

Fund grant 34137-AC8, <strong>and</strong> National Science Foundation<br />

grant EAR-9725308 to J.F. Read, <strong>and</strong> funding from Mobil Oil with<br />

the strong support of Ron Kreisa <strong>and</strong> Jim Markello. We thank<br />

American Association of Petroleum Geologists, Geological Society<br />

of America, <strong>and</strong> the Society of Professional Well Log Analysts for<br />

additional support. Dave Hunt, Steve Dorobek, <strong>and</strong> Wayne Ahr<br />

greatly improved the paper with their insightful reviews.<br />

REFERENCES<br />

AHR, W.M., 1973, The <strong>carbonate</strong> ramp—an alternative to the shelf model:<br />

Gulf Coast Association of Geological Societies, Transactions, v. 23, p.<br />

221–225.<br />

AL-TAWIL, A., 1998, High resolution sequence stratigraphy of Late Mississippian<br />

<strong>carbonate</strong>s in the Appalachian Basin: Unpublished Ph.D.<br />

dissertation, Virginia Tech, Blacksburg, 109 p.<br />

ARKLE, T., JR., BEISSELL, D.R., LARESE, R.E., NUHFER, E.B., PATCHEN, D.G.,<br />

SMOSNA, R.A., GILLESPIE, W.H., LUND, R., NORTON, W., AND PFEFFERKORN,<br />

H.W., 1979, The Mississippian <strong>and</strong> Pennsylvanian (Carboniferous)<br />

Systems in the United States—West Virginia <strong>and</strong> Maryl<strong>and</strong>: U.S.<br />

Geological Society, Professional Paper 1110-D, 35 p.<br />

BARTLETT, C.S., 1974, Anatomy of the Lower Mississippian Delta in southwestern<br />

Virginia: Unpublished Ph.D. dissertation, University of Tennessee,<br />

Knoxville, 373 p.<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

31<br />

BARTLETT, C.S., JR., AND WEBB, H.W., 1971, Geology of the Bristol <strong>and</strong><br />

Wallace quadrangles, Virginia: Virginia Division of Mineral Resources,<br />

Report of Investigations 25, 93 p.<br />

BAXTER, J.W., AND BRENCKLER, P.L., 1982, Preliminary statement on Mississippian<br />

calcareous foraminiferal successions of the midcontinent<br />

(U.S.A.) <strong>and</strong> their correlation to western Europe: Newsletters on<br />

Stratigraphy, v. 11, p. 136–153.<br />

BJERSTEDT, T.W., AND KAMMER, T.W., 1988, Genetic stratigraphy <strong>and</strong> depositional<br />

systems of the upper Devonian–lower Mississippian Price–<br />

Rockwell deltaic complex in the central Appalachians, U.S.A.: Sedimentary<br />

Geology, v. 54, p. 265–301.<br />

BRANSON, E.B., 1912, A Mississippian delta: Geological Society of America,<br />

Bulletin, v. 23, p. 447–456.<br />

BREZINSKI, D.K., 1989, Upper Mississippian depositional patterns in the<br />

north-central Appalachian Basin, <strong>and</strong> their implications to<br />

Chesterian hierarchical stratigraphy: Southeastern Geology, v. 30,<br />

p. 1–23.<br />

BUTTS, C., 1922, The Mississippian Series of eastern Kentucky: Kentucky<br />

Geological Survey, ser. 6, v. 7, 188 p.<br />

BUTTS, C., 1940, Geology of the Appalachian Valley in Virginia (Part I—<br />

Geologic Text <strong>and</strong> Illustrations): Virginia Geological Survey, Bulletin<br />

52, p. 1–568.<br />

BUTTS, C., 1941, Geology of the Appalachian Valley in Virginia (Part II—<br />

Fossil Plates <strong>and</strong> Explanations): Virginia Geological Survey, Bulletin<br />

52, p. 1–568.<br />

CARNEY, C., 1993, The drowning of ooid shoals: Mississippian Greenbrier<br />

Limestone near the West Virginia Dome, in Keith, B.D., <strong>and</strong> Zuppman,<br />

C.W., eds., Mississippian Oolites <strong>and</strong> Modern Analogs: American<br />

Association of Petroleum Geologists, Studies in Geology no. 35, p.<br />

141–148.<br />

CARNEY, C., AND SMOSNA, R., 1989, Carbonate deposition in a shallow<br />

marine gulf, the Mississippian Greenbrier Limestone of the Central<br />

Appalachian Basin: Southeastern Geology, v. 30, p. 25–48.<br />

CAUDILL, M.R., DRIESE, S.G., AND MORA, C.I., 1996, Preservation of a paleo-<br />

Vertisol <strong>and</strong> an estimate of late Mississippian paleoprecipitation:<br />

Journal of Sedimentary Research, v. 66, p. 58–70<br />

COLLINSON, C.W., REXROAD, C.B., AND THOMPSON, T.L., 1971, Conodont<br />

zonation of the North American Mississippian, in Sweet, W.C., <strong>and</strong><br />

Bergstrom, S.M., eds., Symposium on Conodont Biostratigraphy:<br />

Geological Society of America, Memoir 127, p. 353–394.<br />

COLTON, G.W., 1970, The Valley <strong>and</strong> Ridge <strong>and</strong> Appalachian Plateau;<br />

stratigraphy <strong>and</strong> sedimentation; the Appalachian Basin; its depositional<br />

sequences <strong>and</strong> their geologic relationships, in Fisher, G.W.,<br />

Pettijohn, F.J., Reed, J.C., Jr., <strong>and</strong> Weaver, K.N., eds., Studies of<br />

Appalachian Geology, Central <strong>and</strong> Southern: Wiley-Interscience, p.<br />

5–47.<br />

COOPER, B.N., 1948, Status of Mississippian stratigraphy in the central <strong>and</strong><br />

northern Appalachian region, in Weller, J.M., ed., Symposium on<br />

Problems of Mississippian Stratigraphy <strong>and</strong> Correlation: Journal of<br />

Geology, v. 56, p. 255–263.<br />

CROWELL, J.C., 1978, Gondwanan glaciation, cyclothems, continental positioning,<br />

<strong>and</strong> climate change: American Journal of Science, v. 278, p.<br />

1345–1378.<br />

DE WITT, W., AND MCGREW, L.W., 1979, The Appalachian Basin, in Craig,<br />

L.C., <strong>and</strong> Connor, C.W., coordinators, Paleotectonic Investigations of<br />

the Mississippian System in the United States: U.S. Geological Survey,<br />

Professional Paper 1010, part I, p. 59–106.<br />

DENNISON, J.M., 1985, Catskill delta shallow marine strata, in Woodrow,<br />

D.L., <strong>and</strong> Sevon, W.D., eds., The Catskill Delta: Geological Society of<br />

America, Special Paper 201, p. 91–106.<br />

DEVER, G.R., JR., 1973, Stratigraphic relationships in the lower <strong>and</strong><br />

middle Newman Limestone (Mississippian), east-central <strong>and</strong> northeastern<br />

Kentucky: Unpublished M.S. Thesis, University of Kentucky,<br />

Lexington, 121 p.


32<br />

DEVER, G.R., JR., 1986, Mississippian reactivation along the Irvine–Paint<br />

Creek Fault System in the Room Trough, east central Kentucky:<br />

Southeastern Geology, v. 27, no. 2, p. 8–18.<br />

DEVER, G.R., 1995, Tectonic implications of erosional <strong>and</strong> depositional<br />

features in upper Meramecian <strong>and</strong> lower Chesterian (Mississippian)<br />

<strong>carbonate</strong> rocks of south-central <strong>and</strong> east-central Kentucky:<br />

Unpublished Ph.D. Dissertation, University of Kentucky, Lexington,<br />

152 p.<br />

DEVER, G.R., GREB, S.F., MOODY, J.R., CHESTNUT, D.R., KEPFERLE, R.C., AND<br />

SERGEANT, R.E., 1990, Tectonic implications of depositional <strong>and</strong><br />

erosional features in Carboniferous rocks of south-central Kentucky:<br />

Field Guide for Annual Field Conference of the Geological<br />

Society of Kentucky, 53 p.<br />

DIRENZO, V.N., JR., 1986, Petrology <strong>and</strong> depositional environments of<br />

the Little Valley Limestone (Upper Mississippian) Washington<br />

County, Virginia: Unpublished M.S. Thesis, East Carolina University,<br />

Greenville, 154 p.<br />

DODD, J.R., ZUPPMAN, C.W., HARRIS, C.D., LEONARD, K.W., AND BROWN,<br />

T.W., 1993, Petrologic method for distinguishing eolian <strong>and</strong> marine<br />

grainstones, Ste. Genevieve Limestone (Mississippian) of Indiana,<br />

in Keith, B.D., <strong>and</strong> Zuppman, C.W., eds., Mississippian Oolites <strong>and</strong><br />

Modern Analogs: American Association of Petroleum Geologists,<br />

Studies in Geology no. 35, p. 49–59.<br />

ETTENSOHN, F.R., 1975, Stratigraphic <strong>and</strong> paleoenvironmental aspects of<br />

Upper Mississippian rocks (Upper Newman Group), east-central<br />

Kentucky: Unpublished Ph.D. Thesis, University of Illinois, Urbana,<br />

320 p.<br />

ETTENSOHN, F.R., 1980, An alternative to the barrier-shoreline model for<br />

deposition of Mississippian <strong>and</strong> Pennsylvanian rocks in northeastern<br />

Kentucky: Geological Society of America, Bulletin, v. 91, pt. 1, p.<br />

130–135; pt. 2, p. 934–1056.<br />

ETTENSOHN, F.R., 1994, Tectonic control on formation <strong>and</strong> cyclicity of<br />

major Appalachian unconformities <strong>and</strong> associated stratigraphic<br />

sequences, in Dennison, J.M., <strong>and</strong> Ettensohn, F.R., eds., Tectonic <strong>and</strong><br />

Eustatic Controls on Sedimentary Cycles: SEPM, Concepts in Sedimentology<br />

<strong>and</strong> Paleontology, no. 4, p. 217–242.<br />

ETTENSOHN, F.R., DEVER, G.R., AND GROW, J.S., 1988, A paleosol interpretation<br />

for profiles exhibiting subaerial exposure crusts from the<br />

Mississippian of the Appalachian basin, in Reinhart, J., <strong>and</strong> Sigleo,<br />

W.R., eds., Paleosols <strong>and</strong> Weathering Through Geologic Time; Principles<br />

<strong>and</strong> Applications: Geological Society of America, Special<br />

Paper 216, p. 49–79.<br />

ETTENSOHN, F.R., RICE, C.L., DEVER, G.R., AND CHESTNUT, D.R., 1984, Slade<br />

<strong>and</strong> Paragon Formations—new stratigraphic nomenclature for Mississippian<br />

rocks along the Cumberl<strong>and</strong> Escarpment in Kentucky:<br />

U.S. Geological Survey, Bulletin 1605-B, 37 p.<br />

FISCHER, A.G., 1982, Long-term climatic oscillations recorded in stratigraphy,<br />

in climate in Earth history: Washington, D.C., National<br />

Academy Press, p. 97–104.<br />

FLOWERS, R.R., 1956, A regional subsurface study of the Greenbrier<br />

Limestone in West Virginia: Unpublished M.S. Thesis, University of<br />

West Virginia, Morgantown, 220 p.<br />

FRAKES, L.A., FRANCIS, J.E., AND SYKTUS, J.L., 1992, Climate Modes of<br />

the Phanerozoic: Cambridge, U.K., Cambridge University Press,<br />

274 p.<br />

GOLDHAMMER, R., DUNN, K., AND HARDIE, L.A., 1990, Depositional cycles,<br />

composite sea level changes, cycle stacking patterns, <strong>and</strong> the hierarchy<br />

of stratigraphic forcing: examples from Alpine Triassic platform<br />

<strong>carbonate</strong>s: Geological Society of America, Bulletin, v. 102, p. 535–<br />

562.<br />

GOLDHAMMER, R.K., OSWALD, E.J., AND DUNN, P.A., 1991, Hierarchy of<br />

stratigraphic forcing: Example from Middle Pennsylvanian shelf<br />

<strong>carbonate</strong>s of the Paradox basin, in Franseen, E.K., Watney, W.L.,<br />

Kendall, C.G.St.C., <strong>and</strong> Ross, W., eds., Sedimentary Modeling; Com-<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

puter Simulations <strong>and</strong> Methods for Improved Parameter Definition:<br />

Kansas Geological Survey, Bulletin 233, p. 361–414.<br />

GOLONKA, J., ROSS, M.I., AND SCOTESE, C.R., 1994, Phanerozoic paleogeographic<br />

<strong>and</strong> paleoclimatic modeling maps, in Embry, A.F.,<br />

Beauchamp, B., <strong>and</strong> Glass, D.J., eds., Pangea; Gobal Environments<br />

<strong>and</strong> Resources: Canadian Society of Petroleum Geologists, Memoir<br />

17, p. 1–47.<br />

HARRISON, R.S., AND STEINEN, R.P., 1978, Subaerial crusts, caliche profiles,<br />

<strong>and</strong> breccia horizons: comparison of some Holocene <strong>and</strong> Mississippian<br />

exposure surfaces, Barbados <strong>and</strong> Kentucky: Geological Society<br />

of America, Bulletin, v. 89, p. 385–396.<br />

HECKEL, P.H., 1986, Sea-level curve for Pennsylvanian eustatic marine<br />

transgressive–regressive depositional cycles along the midcontinent<br />

outcrop belt, North America: Geology, v. 14, p. 330–334.<br />

HORBURY, A.D., 1989, The relative roles of tectonism <strong>and</strong> eustasy in the<br />

deposition of the Urswick Limestone in south Cumbria <strong>and</strong> north<br />

Lancashire, in Arthurton, R.S., Gutteridge, P., <strong>and</strong> Nolan, S.C., eds.,<br />

The Role of Tectonics in Devonian <strong>and</strong> Carboniferous Sedimentation<br />

in the British Isles: Yorkshire Geological Society, Occasional<br />

Publication no. 6, p. 153–169.<br />

HUGGINS, M.J., 1983, Meramecian conodonts <strong>and</strong> biostratigraphy of the<br />

(Upper Mississippian) Greenbrier Limestone (Hurricane Ridge <strong>and</strong><br />

Greendale synclines), south-western Virginia <strong>and</strong> southern West<br />

Virginia: Unpublished Ph.D. Dissertation, Virginia Tech, Blacksburg,<br />

303 p.<br />

HUNTER, R.E., 1993, An eolian facies in the Ste. Genevieve Limestone of<br />

southern Indiana, in Keith, B.D., <strong>and</strong> Zuppman, C.W., eds., Mississippian<br />

Oolites <strong>and</strong> Modern Analogs: American Association of<br />

Petroleum Geologists, Studies in Geology no. 35, p. 31–48.<br />

IMBRIE, J., AND IMBRIE, K.P., 1986, Ice Ages; Solving the Mystery: Cambridge,<br />

Massachusetts, Harvard University Press, 224 p.<br />

KEITH, B.D., AND ZUPPMAN, C.W., 1990, eds., Mississippian Oolites <strong>and</strong><br />

Modern Analogs: American Association of Petroleum Geologists,<br />

Studies in Geology no. 35, 265 p.<br />

KELLEHER, G.T., AND SMOSNA, R., 1993, Oolitic tidal-bar reservoirs in the<br />

Mississippian Greenbrier Group of West Virginia, in Keith, B.D.,<br />

<strong>and</strong> Zuppman, C.W., eds., Mississippian Oolites <strong>and</strong> Modern Analogs:<br />

American Association of Petroleum Geologists, Studies in<br />

Geology no. 35, p. 163–173.<br />

KHETANI, A.B., AND READ, J.F., 2002, Sequence development of a mixed<br />

<strong>carbonate</strong>–siliciclastic high-relief ramp, Mississippian, Kentucky,<br />

U.S.A.: Journal of Sedimentary Research, v. 72, p. 657–672.<br />

KOERSCHNER, W.F., III, AND READ, J.F., 1989, Field <strong>and</strong> modelling studies<br />

of Cambrian <strong>carbonate</strong> cycles, Virginia Appalachians: Journal of<br />

Sedimentary Petrology, v. 59, p. 654–687.<br />

LEONARD, A.D., 1968, The petrology <strong>and</strong> stratigraphy of Upper Mississippian<br />

Greenbrier Limestones of eastern West Virginia: Unpublished<br />

Ph.D. Thesis, University of West Virginia, Morgantown,<br />

220 p.<br />

LEETARU, H.E., 2000, Sequence stratigraphy of the Aux Vases S<strong>and</strong>stone:<br />

A major oil producer in the Illinois Basin: American Association of<br />

Petroleum Geologists, Bulletin, v. 84, p. 399–422.<br />

MACQUOWN, W.C., AND PEAR, J.L., 1983, Regional <strong>and</strong> local geologic<br />

factors control Big Lime stratigraphy <strong>and</strong> exploration for petroleum<br />

in eastern Kentucky, in Luther, M.K., ed., Proceedings of the technical<br />

sessions, Kentucky Oil <strong>and</strong> Gas Association, 44th Annual Meeting,<br />

June 11–13, 1980: Kentucky Geological Survey, ser. 11, Special<br />

Publication 9, p. 1–20.<br />

MAPLES, C.G., AND WATERS, J.A., 1987, Redefinition of the Meramecian/<br />

Chesterian Boundary (Mississippian): Geology, v. 15, p. 647–651.<br />

MCFARLAN, A.C., SWANN, D.H., WALKER, F.H., AND NOSOW, E., 1955, Some<br />

old Chester problems—correlations of lower <strong>and</strong> middle Chester<br />

formations of western Kentucky: Kentucky Geological Survey Series<br />

IX, Bulletin 16, 37 p.


MCFARLAN, A.C., AND WALKER, F.H., 1956, Some old Chester problems—<br />

correlations along the eastern belt of outcrop: Kentucky Geological<br />

Survey, series IX, Bulletin 20, 36 p.<br />

MERKELY, P.A., 1991, Origin <strong>and</strong> distribution of <strong>carbonate</strong> eolianites in<br />

the Ste. Genevieve Limestone (Mississippian) of southern Indiana<br />

<strong>and</strong> northwestern Kentucky: Unpublished M.S. Thesis, Indiana<br />

University, Bloomington, 147 p.<br />

MII, H., GROSSMAN, E.L., AND YANCEY, T., 1999, Carboniferous isotope<br />

stratigraphies of North America: Implications for Carboniferous<br />

paleoceanography <strong>and</strong> Mississippian glaciation: Geological Society<br />

of America, Bulletin, v. 111, p. 960–973.<br />

MILLER, D.J., AND ERIKSSON, K.A., 1997, Late Mississippian prodeltaic<br />

rhythmites in the Appalachian Basin: A hierarchical record of tidal<br />

<strong>and</strong> climatic periodicities: Journal of Sedimentary Research, v. 67, p.<br />

653–660.<br />

MILLER, D.J., AND ERIKSSON, K.A., 1999, Linked sequence development<br />

<strong>and</strong> global climate change; the Upper Mississippian record in the<br />

Appalachian Basin: Geology, v. 27, p. 35–38.<br />

MITCHUM, R.M., JR., AND VAN WAGONER, J.C., 1991, High frequency<br />

sequences <strong>and</strong> their stacking patterns: sequence-stratigraphic evidence<br />

of high-frequency eustatic cycles: Sedimentary Geology, v.<br />

70, p. 131–160.<br />

NIEMANN, J.C., AND READ, J.F., 1988, Regional cementation from unconformity-recharged<br />

aquifer <strong>and</strong> burial fluids, Mississippian Newman<br />

Limestone, Kentucky: Journal of Sedimentary Petrology, v. 58, p.<br />

688–705.<br />

PRYOR, W.A., AND SABLE, E.G., 1974, Carboniferous of the Eastern Interior<br />

Basin, in Briggs, G., Carboniferous of the Southeastern United<br />

States: Geological Society of America, Special Paper 148, p. 281–313.<br />

PURSER B.H., AND SEIBOLD, E., 1973, The principal environmental factors<br />

influencing Holocene sedimentation <strong>and</strong> diagenesis in the Persian<br />

Gulf, in Purser, B.H., ed., The Persian Gulf: Holocene Carbonate<br />

Sedimentation <strong>and</strong> Diagenesis in a Shallow Epicontinental Sea:<br />

New York, Springer-Verlag, p. 1–10.<br />

QUINLAN, G.M., AND BEAUMONT, C., 1984, Appalachian thrusting, lithospheric<br />

flexure <strong>and</strong> Paleozoic stratigraphy of the eastern interior of<br />

North America: Canadian Journal of Earth Sciences, v. 21, p. 973–976.<br />

READ, J.F., 1985, Carbonate platform facies models: American Association<br />

of Petroleum Geologists, Bulletin, v. 69, p. 1–21.<br />

READ, J.F., 1995, Overview of <strong>carbonate</strong> platform sequences, cycle stratigraphy<br />

<strong>and</strong> reservoirs in greenhouse <strong>and</strong> icehouse worlds, in Read,<br />

J.F., Kerans, C., Weber, L.J., Sarg, J.F., <strong>and</strong> Wright F.W., Milankovitch<br />

Sea Level Changes, Cycles <strong>and</strong> Reservoirs on Carbonate Platforms<br />

in Greenhouse <strong>and</strong> Icehouse Worlds: SEPM, Short Course Notes no.<br />

35, p. 1–102.<br />

REGER, D.B., 1926, Mercer, Monroe, <strong>and</strong> Summer Counties: West Virginia<br />

Geological Survey, 963 p.<br />

RICE, C.L., SABLE, E.G., DEVER, G.R., JR., AND KEHN, T.M., 1979, The<br />

Mississippian <strong>and</strong> Pennsylvanian (Carboniferous) Systems in the<br />

United States—Kentucky: U.S. Geological Survey, Professional Paper<br />

1110-F, 32 p.<br />

REXROAD, C., AND CLARKE, C.E., 1960, Conodonts form the Glen Dean<br />

Formation of Kentucky <strong>and</strong> equivalent formations of Virginia <strong>and</strong><br />

West Virginia: Journal of Paleontology, v. 34, p. 1202–1206.<br />

ROBERTS, J., CLAOUE-LONG, J., JONES, P.J., AND FOSTER, C.B., 1995, SHRIMP<br />

zircon age control of Gondwana sequences in Late Carboniferous<br />

<strong>and</strong> early Permian Australia, in Non-Biostratigraphical Methods of<br />

Dating <strong>and</strong> Correlation: Geological Society of London, Special Publication<br />

89, p. 145–174.<br />

ROSS, C.A., AND ROSS, J.R., 1987, Late Paleozoic sea-levels <strong>and</strong> depositional<br />

sequences, in Ross, C.A., <strong>and</strong> Haman, D., eds., Timing <strong>and</strong><br />

Depositional History of Eustatic Sequences: Constraints on Seismic<br />

Stratigraphy: Cushman Foundation for Foraminiferal Research,<br />

Special Publication 24, p. 137–149.<br />

MISSISSIPPIAN RAMP, WEST VIRGINIA, U.S.A.<br />

33<br />

RUDDIMAN, W.F., RAYMO, M., AND MACINTYRE, A., 1986, Matuyuma 41,000year<br />

cycles: North Atlantic Ocean <strong>and</strong> Northern Hemisphere icesheets:<br />

Earth <strong>and</strong> Planetary Science Letters, v. 80, p. 117–129.<br />

SABLE, E.G., AND DEVER, G.R., 1990, Mississippian Rocks in Kentucky: U.S.<br />

Geological Survey, Professional Paper 1503, 125 p.<br />

SANDO, W.J., 1984, Revised Mississippian time scale, Western Interior<br />

region, conterminous United States: U.S. Geological Survey, Stratigraphic<br />

Notes, p. A15-A26.<br />

SARG, J.F., 1988, Carbonate sequence stratigraphy, in Wilgus, C.K., Hastings,<br />

B.S., Kendall, C.G.St.C, Posamentier, H.W., Ross, C.A., <strong>and</strong> Van<br />

Wagoner, J.C., Sea-Level Changes: An Integrated Approach: SEPM,<br />

Special Publication 42, p. 155–181.<br />

SCOTESE, C.R., AND MCKERROW, W.S., 1990, Revised world maps <strong>and</strong><br />

introduction, in McKerrow, W.S., <strong>and</strong> Scotese, C.R., eds., Palaeozoic<br />

Palaeogeography <strong>and</strong> Biogeography: Geological Society of London,<br />

Memoir 12, p. 1–24.<br />

SHUMAKER, R.C., AND WILSON, T.H., 1996, Basement structure of the Appalachian<br />

forel<strong>and</strong> in West Virginia: Its style <strong>and</strong> effect on sedimentation,<br />

in van der Pluijm, B.A., <strong>and</strong> Catacosinos, P.A., eds., Basement<br />

<strong>and</strong> Basins of Eastern North America: Geological Society of America,<br />

Special Paper 308, p. 139–155.<br />

SMITH, L.B., 1996, High resolution sequence stratigraphy of late Mississippian<br />

(Chesterian) mixed <strong>carbonate</strong>s <strong>and</strong> siliciclastics, Illinois Basin:<br />

Unpublished Ph.D. Thesis, Virginia Tech, Blacksburg, 146 p.<br />

SMITH, L.B., AND READ, J.F., 1999, Application of high resolution sequence<br />

stratigraphy to tidally influenced Late Mississippian <strong>carbonate</strong>s,<br />

Illinois Basin, in Harris, P.M., Saller, A.H., <strong>and</strong> Simo, J.A., eds.,<br />

Advances in Carbonate Sequence Stratigraphy—Application to Reservoirs,<br />

Outcrops, <strong>and</strong> Models: SEPM, Special Publication 63, p. 107–<br />

126.<br />

SMITH, L.B., JR., AND READ, J.F., 2000, Rapid onset of Late Paleozoic glaciation:<br />

Evidence from Upper Mississippian strata of the midcontinent,<br />

United States: Geology, v. 28, p. 279–282.<br />

SMITH, L.B., AND READ, J.F., 2001, Discrimination of local <strong>and</strong> global effects<br />

on Upper Mississippian stratigraphy, Illinois Basin, U.S.A.: Journal of<br />

Sedimentary Research, v. 71, p. 985–1002.<br />

SMITH, T.M., AND DOROBEK, S.L., 1993, Stable isotopic composition of<br />

meteoric calcites: evidence of Early Mississippian climate change in<br />

the Mission Canyon Formation, Montana: Tectonophysics, v. 222, p.<br />

317–331.<br />

SMOSNA, R., AND KOEHLER, B., 1993, Tidal origin of a Mississippian oolite on<br />

the West Virginia Dome, in Keith, B.D., <strong>and</strong> Zuppman, C.W., eds.,<br />

Mississippian Oolites <strong>and</strong> Modern Analogs: American Association of<br />

Petroleum Geologists, Studies in Geology no. 35, p. 149–162.<br />

STAMM, R., 1997, Late Mississippian conodont biostratigraphy of the<br />

Appalachian Basin: Preliminary correlations to the Eastern Interior<br />

Basin <strong>and</strong> eustatic curves (abstract): The Geological Society of America,<br />

31 st Annual South-Central <strong>and</strong> 50 th Rocky Mountain Sections, Abstracts<br />

with Programs, p. 48.<br />

SWANN, D.H., 1964, Late Mississippian rhythmic sediments of Mississippi<br />

Valley: American Association of Petroleum Geologists, Bulletin, v. 48,<br />

p. 637–658.<br />

THOMAS, W.A., 1977, Evolution of Appalachian salients <strong>and</strong> recesses from<br />

reentrants <strong>and</strong> promontories in the continental margin: American<br />

Journal of Science, v. 277, p. 1233–1278.<br />

VEEVERS, J.J., AND POWELL, C.M., 1987, Late Paleozoic glacial episodes in<br />

Gondwanal<strong>and</strong> reflected in transgressive–regressive sequences in<br />

Euramerica: Geological Society of America, Bulletin, v. 98, p. 475–<br />

487.<br />

WEBER, L.J., SARG, J.F., AND WRIGHT, F.M., 1995, Sequence stratigraphy<br />

<strong>and</strong> reservoir delineation of the Middle Pennsylvanian<br />

(Desmoinesian), Paradox basin <strong>and</strong> Aneth Field, southwestern USA,<br />

in Read, J.F., Kerans, C., Weber, L.J., Sarg, J.F., <strong>and</strong> Wright, F.M.,<br />

eds., Milankovitch Sea-Level Changes, Cycles <strong>and</strong> Reservoirs on


34<br />

Carbonate Platforms in Greenhouse <strong>and</strong> Ice-house Worlds: SEPM,<br />

Short Course 35, Part 3, p. 1–81.<br />

WELLS, D., 1950, Lower Middle Mississippian of southeastern West Virginia:<br />

American Association of Petroleum Geologists, Bulletin, v. 34,<br />

p. 882–992.<br />

WHITEHEAD, N.H., III, 1984, Paleogeography <strong>and</strong> depositional environments<br />

of the Lower Mississippian of the East-Central United States, in<br />

Nassichuk, W.W., ed., Paleogeography <strong>and</strong> Paleotectonics: Neuvième<br />

Congrès International de Stratigraphie et de Géologie du Carbonifère<br />

1979, v. 3, Part 2, p. 280–290.<br />

AUS AL-TAWIL, THOMAS C. WYNN, AND J. FRED READ<br />

APPENDIX. —Sections used in the study.<br />

WRIGHT, V.P., AND VANSTONE, S.D., 2001, Onset of Late Paleozoic glacioeustasy<br />

<strong>and</strong> the evolving climates of low latitude areas: A synthesis of<br />

current underst<strong>and</strong>ing: Geological Society of London, Journal, v. 158,<br />

p. 579–582.<br />

YEILDING, C.A., AND DENNISON, J.M., 1986, Sedimentary response to Mississippian<br />

tectonic activity at the east end of the 38th Parallel fracture<br />

zone: Geology, v. 14, p. 621–624.<br />

Section # Name County, State Latitude Longitude<br />

1 Canaan Valley Tucker County, West Virginia 39.002800 -79.467500<br />

2 R<strong>and</strong>olph 28 R<strong>and</strong>olph County, West Virginia 38.716940 -80.132780<br />

3 Mingo R<strong>and</strong>olph County, West Virginia 38.488200 -80.041700<br />

4 Greenbrier 21 Greenbrier County, West Virginia 37.965000 -80.627220<br />

5 Summers 7 Summers County, West Virginia 37.794170 -80.744170<br />

6 Sunoil Test Core #2 Pocahontas County, West Virginia 38.108333 -80.266667<br />

7 Greenbrier 15 Greenbrier County, West Virginia 37.751390 -80.586670<br />

8 Union Monroe County, West Virginia 37.610800 -80.587500<br />

9 Mercer 25 Mercer County, West Virginia 37.440000 -81.018610<br />

10 Ingleside Mercer County, West Virginia 37.313900 -81.141700<br />

11 Bristol Washington County, Virginia 36.75 -82.05<br />

12 Wayne 277 Wayne County, West Virginia 38.271111 -82.538611<br />

13 Wayne 1497 Wayne County, West Virginia 38.1975 -82.334167<br />

14 Lincoln 155 Lincoln County, West Virginia 38.306111 -82.202222<br />

15 Lincoln 1375 Lincoln County, West Virginia 38.163060 -82.130280<br />

16 Mingo 796 Mingo County, West Virginia 37.924170 -82.206670<br />

17 Logan 102 Logan County, West Virginia 37.96861 -81.973056<br />

18 Logan 922 Logan County, West Virginia 37.835280 -81.960000<br />

19 Logan 388 Logan County, West Virginia 37.78972 -81.784444<br />

20 Wyoming 253 Wyoming County, West Virginia 37.77167 -81.51806<br />

21 Wyoming 87 Wyoming County, West Virginia 37.525556 -81.637500<br />

22 McDowell 63 McDowell County, West Virginia 37.357500 -81.653611<br />

23 McDowell 792 McDowell County, West Virginia 37.44306 -81.400833<br />

24 Mercer 16 Mercer County, West Virginia 37.50833 -81.17639<br />

25 Mercer 1 Mercer County, West Virginia 37.328056 -81.224444<br />

26 Mercer 27 Mercer County, West Virginia 37.39722 -81.0525


Personal Information:<br />

Society for Sedimentary Geology<br />

Membership Application<br />

Last Name, First (Given) Middle Initial Date of Birth<br />

Preferred Mailing Address<br />

City State/Province Postal Code<br />

Country Phone Number<br />

Fax Number Citizenship E-mail Address<br />

Academic Record:<br />

From To Institution Major Degree/Date<br />

Professional Record:<br />

From To Employer Address Nature of Work<br />

Affiliations:<br />

AAPG AGU<br />

PS GSA<br />

IAS Other: ___________<br />

membership type:<br />

Sustaining Member ................................................ $300<br />

Financial support may be dedicated to a specific program.<br />

Voting Member .............................. see Journal Choice<br />

Requires 3 years professional experience past bachelors degree.<br />

Associate Member ....................... see Journal Choice<br />

Waives professional experience. May not vote or hold office.<br />

Spouse Member ....................................................... $25<br />

Membership at same address without additional journal.<br />

Payment Information:<br />

References:<br />

1<br />

Name<br />

Address<br />

Name<br />

Address<br />

check/money order enclosed Total Amount:_______________<br />

bill my credit card<br />

Visa American Express MasterCard<br />

Account Number: _________________________________________________<br />

Expiration Date: __________________________________________________<br />

Printed Name: ____________________________________________________<br />

Signature: _______________________________________________________<br />

2<br />

546<br />

Journal Choice:<br />

Journal of Sedimentary Research<br />

Print: $75<br />

PALAIOS<br />

Online: $70 Both: $85<br />

Print: $75<br />

Both Journals<br />

Online: $70 Both: $85<br />

Print: $120 Online: $110 Both: $130<br />

Add $10 surcharge to defray mailing outside the USA<br />

Fax this form with credit<br />

card information to:<br />

918.621.1685<br />

918.621.1685<br />

or mail with payment to:<br />

SEPM SEPM Membership<br />

Membership<br />

6128 6128 E. E. 38th 38th Street, Street, #308<br />

#308<br />

Tulsa, ulsa, OK OK OK 74135, 74135, USA<br />

USA

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!