stratigraphic analysis of the - university of texas at austin

66

Upload: others

Post on 27-Oct-2021

5 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Stratigraphic Analysis of the - University of Texas at Austin
Page 2: Stratigraphic Analysis of the - University of Texas at Austin

Report of Investigations No. 201

Stratigraphic Analysis of theUpper Devonian Woodford Formation,

Permian Basin, West Texas andSoutheastern New Mexico

John B. Comer*

*Current addressIndiana Geological Survey

Bloomington, Indiana 47405

1991

Bureau of Economic Geology • W. L. Fisher, DirectorThe University of Texas at Austin • Austin, Texas 78713-7508

Page 3: Stratigraphic Analysis of the - University of Texas at Austin

Contents

Abstract . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .1Introduction .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1Methods .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .3Stratigraphy .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .5

Nomenclature ...................................................................................................................5Age and Correlation ........................................................................................................6Previous Work .................................................................................................................6

Western Outcrop Belt ......................................................................................................6Central Texas ...................................................................................................................7Northeastern Oklahoma and Northern Arkansas .................................................................8Ouachita Fold Belt .......................................................................................................................8Central and Southern Oklahoma ........................................................................................8Permian Basin ..................................................................................................................9

Formation Boundaries ....................................................................................................9Distribution .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .10

Northwestern Shelf and Matador Uplift ....................................................................10Eastern Shelf ...................................................................................................................10Central Basin Platform and Pecos Arch .....................................................................11Delaware Basin ...............................................................................................................11Midland Basin ................................................................................................................11Val Verde Basin ..............................................................................................................11Diablo Platform and Western Outcrop Belt ...............................................................12

Lithofacies .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .12Black Shale ......................................................................................................................12

Characteristic Features ....................................................................................................12Bedding and Sedimentary Structures ...............................................................................12Texture ..........................................................................................................................13Composition ...................................................................................................................13

Siltstone ...........................................................................................................................16Characteristic Features ....................................................................................................16Bedding and Sedimentary Structures ...............................................................................17Texture ..........................................................................................................................17Composition ................................................................................................................... 17

Formation-Boundary Lithologies ................................................................................20Lower Contact ................................................................................................................20Upper Contact ................................................................................................................20

Lithofacies Correlation ..................................................................................................22Lithofacies Distribution ................................................................................................23Depositional Processes ..................................................................................................25

Siltstone .........................................................................................................................25Black Shale .....................................................................................................................26

Lithologic Patterns and Origin of Sediments ............................................................26Depositional Setting ... . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .30

Paleogeography .............................................................................................................30Paleotectonics .................................................................................................................31Paleoclimate ....................................................................................................................32Paleoceanography ..........................................................................................................33

iii

Page 4: Stratigraphic Analysis of the - University of Texas at Austin

Depositional Mechanisms ............................................................................................34Synopsis of Depositional History ................................................................................35

Petroleum Potential . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .36Summary .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .36Acknowledgments .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .38References .. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .39Appendices

A. Location of cores and measured sections ...........................................................................43B. Description of cores and measured sections ......................................................................44C. Point-count data for the Woodford Formation .................................................................60D. Organic content of the Woodford Formation ....................................................................61

Figures1. Index maps showing structural provinces in the Permian Basin ......................................................................22. Map showing lines of cross sections and locations of cores and measured sections ......................................43. Correlation chart for Devonian and Mississippian Systems in West Texas

and southeastern New Mexico................................................................................................................................ 74. Photos of Woodford black shale ...........................................................................................................................145. Photos of Woodford Siltstone................................................................................................................................ 186. Photos of Woodford contacts ................................................................................................................................217. Log correlation of Woodford lithofacies..............................................................................................................238. Fence diagram of Upper Devonian units ............................................................................................................249. Regional lithologic variations in Upper Devonian rocks in West Texas and

southeastern New Mexico .....................................................................................................................................2610. Late Devonian paleogeography of West Texas and southeastern New Mexico ...........................................3011. Late Devonian paleogeography of North America............................................................................................3212. Model of Late Devonian circulation during eustatic highstand ......................................................................34

Plates (in pocket)1. Structure map of Woodford Formation, Permian Basin.2. Isopach map of Woodford Formation, Permian Basin.3. West-east cross section A-A'. Line of section in figure 2.4. West-east cross section B-B'. Line of section in figure 2.5. West-east cross section C-C'. Line of section in figure 2.6. North-south cross section D-D'. Line of section in figure 2.7. North-south cross section E-E'. Line of section in figure 2.

iv

Page 5: Stratigraphic Analysis of the - University of Texas at Austin

AbstractThe Upper Devonian Woodford Formation is an organic-rich petroleum source

rock that extends throughout West Texas and southeastern New Mexico and cur-rently is generating oil or gas in the subsurface. The Woodford is a potential hydro-carbon reservoir in areas where it is highly fractured; the most favorable drillingtargets are fractured siltstone or chert beds in densely faulted regions such as theCentral Basin Platform, southernmost Midland Basin, and parts of the NorthwesternShelf. Stratigraphic analysis was undertaken to determine how the Woodford wasdeposited and why its petroleum source potential is so great.

The Woodford consists of two lithofacies, black shale and siltstone. Black shale, themost widely distributed rock type, is very radioactive and contains varvelike parallellaminae, abundant pyrite, and high concentrations of marine organic matter. Siltstone,typically a basal facies, in deep basin and proximal shelf settings, exhibits disruptedstratification, graded layers, fine-grained Bouma sequences, and a subequal mixtureof silt-sized quartz and dolomite. Black shale is mostly pelagic and represents ananaerobic biofacies, whereas siltstone is the result of bottom-flow deposition andrepresents a dysaerobic biofacies.

The depositional model developed herein for the Woodford was based on strati-graphic sequence, patterns of onlap, and lithologic variations, together with publishedinformation about global paleogeography, paleoclimate, and eustasy. During the LateDevonian, the Permian Basin was a low-relief region located on the western marginof North America in the arid tropics near 15 degrees south latitude. Worldwidemarine transgression caused flooding of the craton and carried water from a zone ofcoastal upwelling into the expanding epeiric sea. Strong density stratificationdeveloped, due partly to accumulation of hypersaline bottom water that formed locallyin the arid climate. Anaerobic conditions resulted from poor vertical circulation andfrom high oxygen demand, which was caused by the decay of abundant organicmatter produced in the nutrient-rich surface waters. Continuous, slow deposition ofpelagic material was interrupted by episodic, rapid deposition of silt and mud frombottom flows generated during frequent tropical storms.

This report documents the composition, distribution, and structure of the WoodfordFormation in a major hydrocarbon-producing basin. Petrologic and organic geochem-ical data helped explain the origin of the unit and provided information necessary forpredicting potential locations and lithologies of commercial petroleum reservoirs withinthe Woodford. Combining comprehensive stratigraphic, petrologic, and geochemicaldata was useful for developing a depositional and exploration model of Devonianblack shale in West Texas and New Mexico. Similar studies should be conductedelsewhere to enable discovery of unconventional hydrocarbon reserves in black shales.

Keywords: Upper Devonian, Woodford, black shale, siltstone, sourcerocks, unconventional reservoirs, depositional model, paleogeography

IntroductionThis report presents a stratigraphic analysis

of the Woodford Formation (Upper Devonian)in the Permian Basin of West Texas and south-eastern New Mexico (fig. la, b). The study is

part of a larger project undertaken to determinehow and why these enigmatic, organic-richmarine rocks were deposited and to documenttheir petroleum-generation history. The part of

1

Page 6: Stratigraphic Analysis of the - University of Texas at Austin

FIGURE 1. Index maps showing structural provinces in the Permian Basin, (a) Late Paleozoic to Recent. AfterWalper (1977) and Hills (1984). (b) Late Devonian. After Galley (1958) and Wright (1979).

2

Page 7: Stratigraphic Analysis of the - University of Texas at Austin

the project involved in this study entailedmapping, conducting lithologic studies of coresand outcrops, and reconstructing paleogeog-raphy and depositional environments.

The Woodford Formation has long beenrecognized by geologists working in the regionas an important stratigraphic marker because ofits black shale lithology, anomalously highradioactivity, and widespread distribution(Ellison, 1950; Wright, 1979). The organic-richformation typically yields shows of oil fromcuttings and cores and produces a gas responseon mudlogs. The Woodford, acknowledged as aprincipal petroleum source rock in the PermianBasin (Galley, 1958; Jones and Smith, 1965;Horak, 1985), contains some intervals of “oilshale” as well (>10 gal of retortable oil per tonof shale). It is also a low-grade, subeconomicuranium and heavy metal deposit (Swanson,1960; Landis, 1962; Duncan and Swanson, 1965).

The economic potential of black shales hasprompted several studies of shale deposition.Previous publications describe and interpretthe origin of Devonian black shales in the east-ern United States (for example, Cluff, 1980;Ettensohn and Barron, 1981; Broadhead andothers, 1982; Schopf, 1983; Ettensohn and Elam,1985; Pashin and Ettensohn, 1987), but no com-

parable work has been published on equivalentstrata in the southern Midcontinent. Develop-ing a comprehensive depositional model of theWoodford was complicated in that no modernanalog is available for comparison. During theLate Devonian a euxinic sea, in which broadexpanses of marine black shale had beendeposited, occupied most of the Midcontinentof North America. However, virtually no largeeuxinic epeiric seas on stable cratons and pas-sive continental margins adjacent to the openocean exist in the modern world. Why cratoniceuxinic seas developed must be understoodbefore the origin of the Woodford can befully explained. Although global controls, suchas deglaciation and ocean ridge expansion(Heckel and Witzke, 1979; Johnson and others,1985), can account for the worldwide trans-gression in the Late Devonian, regional controlsmust be used to account for the stronglystratified water columns and widespreadbottom anoxia that developed in North America.The depositional model described later hereinshows that it was the unique relationship amonggeography, geomorphology, tectonics, climate,and oceanography that produced the uncommonenvironment and unusual lithology of theWoodford Formation.

MethodsData for the project were obtained from

558 well logs, 13 cores, and 3 measured sections.Well control is plotted on the structure andisopach maps (pls. 1, 2), and the location of cores,measured sections, and cross sections is shownin figure 2. An index of well names and locationsis on open file at the Bureau of EconomicGeology, and the core and measured sectionlocalities are listed in appendix A.

Plates 1 and 2 were contoured using the welldata shown on each map. In areas of poorcontrol, elevation of the Woodford (pl. 1) wasinferred using the Tectonic Map of Texas (Ewing,1991), which was contoured on top of theEllenburger Formation (Lower Ordovician) inWest Texas and on top of the Silurian-Devoniancarbonate section in southeastern New Mexico.

Faults mapped in this report were redrawn fromEwing’s map.

Outcrops in the Hueco, Franklin, and Sacra-mento Mountains were described to comparethese well-studied measured sections in the westwith the poorly known rocks in the subsurface.Outcrops were chosen that had been mappedpreviously and for which paleontological analy-sis had established relative ages.

The Woodford Formation was identified fromwell logs, primarily by high radioactivity on thegamma-ray log (pls. 3 through 7), and by itsstratigraphic position between carbonates.Although other highly radioactive strata lie inthe Permian Basin, the Woodford is the mostlaterally persistent and typically exhibits thestrongest radioactivity anomaly.

3

Page 8: Stratigraphic Analysis of the - University of Texas at Austin
Page 9: Stratigraphic Analysis of the - University of Texas at Austin

The Woodford was more difficult to identifywhere it is overlain by radioactive, fine-grainedcarbonates or shales (for example, pl. 5, logs 9through 14; pl. 6, logs 9 through 13; pl. 7, logs 10through 12 and 16, 17) and where the lower partof the formation is much less radioactive thanthe upper part (for example, pl. 4, logs 9, 11;pl. 7, logs 7, 8). In such sections, the upper andlower contacts were picked from cores, ifavailable, or on sonic, resistivity, and neutronlogs. Typically the Woodford Formation exhibitslow sonic velocity, low resistivity, and lowneutron-induced radiation.

In cores and outcrops, the Woodford andcorrelative formations were identified by theirhigh radioactivity and unique lithology. Aradioactivity profile (counts per second [CPS])was made for each core and outcrop using ahand-held scintillometer (Ettensohn and others,1979). Discrepancies between log depth and coredepth were corrected by comparing the radio-activity profile and the wireline gamma-ray log.Lithologically, Woodford black shale contrastssharply with the light-colored Silurian-Devoniancarbonates below and Mississippian carbonatesabove. Where differences in color and compo-sition are less obvious, continuous, varvelikeparallel laminae and abundant pyrite distinguishthe Woodford.

Petrologic analysis was conducted using slab-bed cores, outcrops, slabbed hand specimens,and thin sections (app. B). Uncovered thin sec-tions were X-rayed to identify the clay mineralsand to distinguish calcite and dolomite, and se-lected thin sections were point counted (app. C).

StratigraphyNomenclature

The name Woodford was first used by Taff(1902) to describe exposures of chert and blackshale along the southern flank of the ArbuckleMountain anticline in Carter County, Oklahoma.Both Woodford Chert and Woodford Shale areestablished as formation names (Keroher andothers, 1966), but the term Woodford Formationalso appears in the literature. In this report,“Woodford Formation” is used because of thewide variety of lithologies that compose the

Woodford lithofacies were correlated in thesubsurface using gamma-ray logs (pls. 3through 7). The two dominant lithofacies, blackshale and siltstone, are readily identified becausesiltstone is markedly less radioactive than blackshale (app. B; C5, C9).

Total organic carbon (TOC), vitrinite reflec-tance (% Ro), and kerogen morphology data(app. D) were valuable in interpreting thesediment provenance, paleogeography, cli-mate, and oceanography during Woodforddeposition. Recognition of pelagic and terri-genous sediment was aided by distinguishingbetween amorphous kerogen, which derivesfrom organic matter of aquatic origin, andstructured kerogen, which is mostly the debrisof land plants (Hunt, 1979; Tissot and Welte,1984). Large concentrations of organic matter,recorded in the rocks as high TOC values,indicate high primary productivity, rapidsedimentation, or anoxic conditions; kerogentype records relative influences of terrigenous,paralic, or marine sources and indirectly reflectsdepositional processes, paleosalinity, paleo-climate, and proximity to land (Byers, 1977;Hunt, 1979; Demaison and Moore, 1980; Tissotand others, 1980; Tissot and Welte, 1984; Stein,1986; Pedersen and Calvert, 1990). Vitrinitereflectance, which records maximum paleo-temperature (Dow, 1977; Hunt, 1979; Tissot andWelte, 1984), allows inferences to be made aboutstructural evolution, thermal events, and burialhistory of the basin during and after the LateDevonian and constrains models of mid-Paleozoic tectonics and paleogeography.

interval in the southern Midcontinent. The mostreadily apparent and dominant lithology is blackshale; however, chert, dolostone, sandstone,siltstone, and light-colored shale are common(Harlton, 1956; Amsden and others, 1967;Amsden, 1975, 1980).

Correlation and nomenclature of Devonianand Mississippian formations are well knownregionally (fig. 3), but within the Permian Basin,stratigraphy and correlation of Silurian,Devonian, and Mississippian strata are poorly

5

Page 10: Stratigraphic Analysis of the - University of Texas at Austin

known. Throughout this report, therefore, car-bonate rocks underlying the Woodford arereferred to as Silurian-Devonian (undifferen-tiated), and those overlying the Woodford arereferred to as Mississippian (undifferentiated),unless faunal or lithologic data indicate a specificsystem, series, or stage.

Age and CorrelationThe Woodford Formation is mostly Late

Devonian (Frasnian-Famennian) in age,although beds of latest Middle Devonian(Givetian) and earliest Mississippian (Kinder-hookian) appear at some localities (Hass andHuddle, 1965; Amsden and others, 1967;Amsden and Klapper, 1972; Amsden, 1975,1980). Ellison (1950) found Late Devonianconodont assemblages but no Mississippianfossils in the Woodford Formation in thePermian Basin, and he documented the cor-relation between the Woodford in the PermianBasin and the Percha Formation in southeasternNew Mexico and West Texas (fig. 3). The LateDevonian age of the Percha and Sly GapFormations (fig. 3) has been established byfaunal analysis (Stevenson, 1945; Laudon andBowsher, 1949).

The Woodford is stratigraphically equivalentto several Devonian black shales in NorthAmerica, including the Antrim Shale in theMichigan Basin, the Chattanooga and OhioShales in the Appalachian Basin, the NewAlbany Shale in the Illinois Basin, the BakkenFormation in the Williston Basin, and theExshaw Formation in the Western Canada Basin(Meissner, 1978; Cluff and others, 1981; Roen,1984; Burrowes and Krause, 1987). Correlativerocks exposed in uplifts in the southern Mid-continent include the Houy Formation in theLlano Uplift of Central Texas; the ChattanoogaShale in the Ozark Uplift of northeasternOklahoma, southern Missouri, and northernArkansas; the middle division of the ArkansasNovaculite in the Ouachita Mountains of south-eastern Oklahoma and west-central Arkansas;the upper part of the Caballos Novaculite in theMarathon region of West Texas; the PerchaFormation in the Hueco and Franklin Mountainsof West Texas; and the Sly Gap Formation inthe Sacramento Mountains of southeastern New

Mexico (King and others, 1945; Stevenson, 1945;Laudon and Bowsher, 1949; Graves, 1952; Cloudand others, 1957; Huffman, 1958; Hass andHuddle, 1965; Amsden and others, 1967).

Previous WorkWestern Outcrop Belt

Throughout the Franklin, Hueco, andSacramento Mountains, Middle and UpperDevonian rocks unconformably overlie massivebeds of the Lower Silurian Fusselman Dolo-mite (King and others, 1945; Stevenson, 1945;Laudon and Bowsher, 1949; LeMone, 1971;Lucia, 1971). In the Franklin and Hueco Moun-tains and at Bishop Cap, New Mexico, theFusselman is overlain by the upper Middle tolower Upper Devonian Canutillo Formation,which is overlain conformably by the UpperDevonian Percha Formation (King and others,1945; Rosado, 1970) (app. B; Pl, P4). TheCanutillo consists of dark cherty and nonchertydolostone (Rosado, 1970), and the Percha isblack, fissile, nonfossiliferous shale (Stevenson,1945). The Canutillo-Percha contact is sharp, andthe lithologic transition abrupt.

In the Sacramento Mountains, the Fusselmanis overlain by the upper Middle to lower UpperDevonian Onate Formation, which is overlainby the lower to middle Upper Devonian Sly GapFormation (Stevenson, 1945; Laudon andBowsher, 1949; Kottlowski, 1963; Rosado, 1970;Bolton and others, 1982). Locally, rocks assignedto the Percha overlie the Onate or the Sly Gap(Pray, 1961; Bolton and others, 1982). The Onate-Sly Gap contact was found to be conformableby Stevenson (1945) but locally eroded anddisconformable by Pray (1961). Kottlowski (1963)suggested that the Onate may be a basal faciesof Sly Gap because the contact is gradationalor only slightly erosional. The Onate consists ofinterbedded gray-brown shale, siltstone, finesandstone, and carbonate (Stevenson, 1945), andthe most common lithology is dolomitic siltstone(Kottlowski, 1963). The Sly Gap is fossiliferousand consists of thinly interbedded, mostly tanto pale-yellow shale, siltstone, and limestone,along with a few dolomitic beds (Stevenson,1945; Rosado, 1970). The Sly Gap is distinguishedfrom the Onate in the field by color; and the Sly

6

Page 11: Stratigraphic Analysis of the - University of Texas at Austin

Gap has more shale and fewer massive, resistantbeds than does the Onate (Stevenson, 1945). Inthe Sacramento Mountains, the Sly Gap grad-ually thins to the east and south and containsmore black shale than do exposures farther west(Stevenson, 1945), reflecting a facies relationshipwith the Percha (Rosado, 1970).

At most localities in the Franklin, Hueco, andSacramento Mountains, the Percha and Sly GapFormations are overlain disconformably byMississippian limestones (King and others, 1945;Rosado, 1970) (app. B; P2, P4). At Bishop Cap,New Mexico (app. B; Pl), and locally in theSacramento Mountains, Upper Devonian rocksare overlain conformably by the KinderhookianCaballero Formation (Rosado, 1970; Bolton andothers, 1982) (fig. 3).

Central TexasIn the Llano region of Central Texas, the Houy

Formation disconformably overlies rocks ofEarly to Middle Devonian and Early Ordovicianage (Cloud and others, 1957). Rocks below theunconformity are carbonates, and most arecherty. In upward succession, the Houy consistsof a lower or basal chert breccia (Ives BrecciaMember), black shale (Doublehorn Shale Mem-ber), and an upper, unnamed phosphatic unit.The Ives Breccia consists mostly of angular frag-ments and unbroken nodules of chert and locallycontains angular blocks of dolostone, all of whichappear to be little-moved lag deposits (Cloudand others, 1957). The Doublehorn Shale isfissile, radioactive, spore-bearing black shale, andthe upper phosphatic unit contains phosphatic

FIGURE 3. Correlation chart for Devonian and Mississippian Systems in West Texas and southeasternNew Mexico. Adapted from Rosado (1970), LeMone (1971), Hoenig (1976), Bolton and others (1982),Lindberg (1983), and Hills (1984).

7

Page 12: Stratigraphic Analysis of the - University of Texas at Austin

debris such as fish bones, pellets, and conodonts.The Houy is predominantly Late Devonian, butlocally the lowermost Houy may be MiddleDevonian and the upper phosphatic unit partlyEarly Mississippian (Kinderhookian) (Cloud andothers, 1957).

The Houy is conformably overlain by theKinderhookian Chappel Limestone (Cloud andothers, 1957). However, the upper Houy has thinbeds, interrupted faunal zones, and intervalscontaining mixed Mississippian and Devonianfossils, all of which make correlation, age, andvertical continuity difficult to determine (Cloudand others, 1957).

Northeastern Oklahoma andNorthern Arkansas

In the Ozark Uplift, the Chattanooga Shalerests disconformably on rocks ranging in agefrom Devonian to Ordovician (Huffman, 1958).The Sylamore Sandstone Member constitutes thelower part of the formation at many localities,and its age is late Middle Devonian to lateKinderhookian (Freeman and Schumacher,1969; Pittenger, 1981). The black shale interval,sometimes called the Noel Shale Member(Huffman and Starke, 1960), is predominantlyLate Devonian but ranges in age from early LateDevonian to Kinderhookian (Amsden andothers, 1967). The Sylamore is submature tosupermature quartzarenite that contains minorphosphate, glauconite, and locally abundantdolomite (Pittenger, 1981). Quartz was reworkedfrom contemporaneous exposures of the MiddleOrdovician Bergen Sandstone (Pittenger, 1981).Locally the basal layer of the Sylamore ischert breccia (Amsden and others, 1967). TheNoel is black, fissile, radioactive shale and isthe most abundant Chattanooga lithology. TheChattanooga is overlain disconformably bylimestones and cherts of the Mississippian BooneGroup. The Boone is predominantly Osageanbut ranges in age from middle Kinderhookianto early Meramecian (Sutherland and Manger,1979).

Ouachita Fold BeltThe middle division of the Arkansas

Novaculite in Oklahoma and Arkansas is fromLate Devonian to Kinderhookian age and repre-sents, at least partly, a lateral facies of the

8

Woodford (Hass, 1951; Amsden and others,1967). Likewise, the upper Caballos Novaculitein West Texas is a Late Devonian (Graves, 1952)lateral facies of the Woodford. Faunal data arescarce and contact relationships problematic, butvertical lithologic continuity suggests that theWoodford-equivalent interval in the novaculiteformations is bounded conformably by theunderlying and overlying beds. The upperCaballos contains mostly white novaculite, andthe middle division of the Arkansas Novaculitecontains interbedded dark-gray and greenish-gray shales and dark-gray novaculite (Hass,1951; Amsden and others, 1967).

Central and Southern OklahomaIn central and southern Oklahoma, the

Woodford Formation rests disconformably onrocks of late Early Devonian to Ordovician age(Amsden, 1975, 1980). The Woodford is mostlyLate Devonian but ranges in age from Givetianto Kinderhookian (Hass and Huddle, 1965;Amsden and others, 1967; Amsden and Klapper,1972; Amsden, 1975, 1980). A basal clastic unit,the Givetian to early Famennian Misener Sand-stone, is present in some areas (Amsden andKlapper, 1972). Woodford black shale is Frasnianto Kinderhookian (Hass and Huddle, 1965;Amsden and others, 1967).

Rocks underlying the Woodford are pre-dominantly carbonates, and some are cherty. Insouthern Oklahoma, the Misener is sandstone,siltstone, green shale, dolostone, or chert breccia,whereas in north-central Oklahoma it is mostlymature quartzarenite containing minor glauco-nite and phosphate and locally abundant dolo-mite (Harlton, 1956; Amsden and others, 1967;Amsden and Klapper, 1972; Amsden, 1980;Francis, 1988). Quartz was derived with littletransport from Middle Ordovician Simpsonsandstone in north-central Oklahoma and fromthe Ouachita province in southern Oklahoma(Amsden and Klapper, 1972). Black shale is themost widespread Woodford lithology. It is fissile,spore-bearing, and highly radioactive, and inthe Arbuckle Mountains it is interbedded withchert (Amsden and others, 1967; Amsden, 1975,1980). Woodford chert is dark and rich inradiolarians and marine organic matter (Comerand Hinch, 1987). The Misener-Woodfordsequence is stratigraphically equivalent to the

Page 13: Stratigraphic Analysis of the - University of Texas at Austin

Sylamore-Chattanooga sequence in the OzarkUplift, and the lower boundary of both se-quences is diachronous, onlapping parts of thenorthern Oklahoma shelf and the Ozark Uplift(Freeman and Schumacher, 1969; Amsden andKlapper, 1972; Amsden, 1980).

In the Arbuckle Mountains, the Woodford isconformably overlain by the Sycamore Forma-tion. The Sycamore consists of poorly fossilif-erous, fine-grained, silty limestone interbeddedwith dark shale, and its age spans the Early toMiddle Mississippian (Kinderhookian toMeramecian) (Ham, 1969). In the subsurface, theWoodford is overlain by Mississippian rocks(mostly limestones), but the stratigraphic rela-tionship is problematic. In basinal regions, evi-dence of unconformity is obscure, although thecontact is probably disconformable (Ham andWilson, 1967; Frezon and Jordan, 1979). Locallythe contact appears to be gradational, and theunconformity, if present, represents only a minorstratigraphic break (Frezon and Jordan, 1979).

Permian BasinIn the Permian Basin, lithologic, electric-

log, and sparse faunal data indicate that theWoodford unconformably overlies rocks rangingin age from Devonian to Ordovician (Lloyd,1949; Ellison, 1950; Peirce, 1962; McGlasson, 1967,Munn, 1971; Hoenig, 1976). The Woodford isoverlain disconformably by Mississippian lime-stone (fig. 3) and locally by rocks as young asEarly Permian (Lloyd, 1949; Wright, 1979). Lloyd(1949) assigned the Mississippian limestonesection in the subsurface to the MeramecianRancheria Formation on the basis of a few fossilsand lithologic similarity to rocks exposed in theSacramento Mountains. Older, Osagean andKinderhookian rocks have not generally beenrecognized in the basin, although Kinderhookianstrata were postulated to exist in a small area ofeastern Chaves, southwestern Roosevelt, andnorthwestern Lea Counties, New Mexico, on thebasis of lithologic similarity to the CaballeroFormation in the Sacramento Mountains (Lloyd,1949).

Ellison (1950) divided the Woodford Forma-tion into three units using radioactivity, logresponse, and lithology. The lower unit wascalcareous and cherty, and it had the lowestradioactivity; the middle unit had the most

9

resinous spores and the highest radioactivity;and the upper unit had few spores andintermediate radioactivity (Ellison, 1950). Themiddle unit was the most widespread, com-prising the black shale lithology characteristicof the Woodford throughout the basin, and thelower unit was the most areally restricted. Acorrelation may exist between Ellison’s lowerWoodford unit and the Ives Breccia Member andbetween the middle Woodford unit and theDoublehorn Shale Member of the Houy For-mation in Central Texas (Wright, 1979).

Formation BoundariesThe lower boundary of the Woodford and its

stratigraphic equivalents represents a majorregional unconformity that extends across thesouthern Midcontinent and records a majorperiod of uplift and erosion that is at least partlyDevonian (Galley, 1958; Amsden and others,1967; Ham and Wilson, 1967; Ham, 1969). Duringthis regressive episode, older Devonian andSilurian strata were removed over broad areasof the Midcontinent, and rocks below the un-conformity became locally deeply eroded (Hamand Wilson, 1967; Ham, 1969; Amsden, 1975).In basinal regions, such as the Anadarko Basin,the unconformity marks the end of earlyPaleozoic shallow-water carbonate sedimenta-tion and the beginning of deep-water carbonateand clastic deposition (Ham, 1969). TheWoodford and correlative formations are di-achronous and represent onlapping sediments(Freeman and Schumacher, 1969; Amsden andKlapper, 1972) deposited during worldwide LateDevonian marine transgression (Johnson andothers, 1985). The coarse sandstone and brecciaoccurring locally above the unconformity are lagdeposits derived from older formations (Cloudand others, 1957; Amsden and others, 1967;Amsden and Klapper, 1972; Amsden, 1975, 1980;Pittenger, 1981), and the black shale representsstrongly reduced mud laid down on the anoxicfloor of an epeiric sea (Ellison, 1950; Wright,1979).

The upper boundary of the Woodford repre-sents only a minor stratigraphic break (Ham andWilson, 1967; Frezon and Jordan, 1979; Click,1979; Mapel and others, 1979). It is discon-formable at some localities (for example, the

Page 14: Stratigraphic Analysis of the - University of Texas at Austin

Ozark Uplift and parts of the western outcropbelt and Oklahoma subsurface) and conformableat others (for example, Central Texas, theArbuckle Mountains, and the Ouachita FoldBelt). The local occurrence and minor strati-graphic expression of disconformities indicate

DistributionDistribution of the Woodford Formation in

the Permian Basin is illustrated in plates 1through 7. The area contoured in plates 1 and 2was not extended to the western outcrop beltbecause of limited well control in Chaves, Eddy,Otero, and Lincoln Counties, New Mexico, andCulberson, Hudspeth, and El Paso Counties,Texas.

Relief on the present-day Woodford surfaceis more than 20,000 ft in the subsurface (pl. 1)and more than 25,000 ft if elevations in thewestern outcrop belt are included. Most of therelief in the basin developed as a result ofdeformation during the late Paleozoic Ouachitaorogeny (Galley, 1958; Muehlberger, 1980),whereas relief in the outcrop belt and DiabloPlatform was strongly influenced by laterLaramide deformation (Muehlberger, 1980).

The Woodford Formation ranges in thicknessfrom 0 to 661 ft (pl. 2) and is thickest in structurallows and thinnest or absent on structural highs.Thicknesses shown on plate 2 were not correctedfor dip and do not everywhere represent truestratigraphic thicknesses. The Woodford is morenearly flat-lying in basin and shelf settingsfarthest from major faults (for example, on theEastern Shelf and in most parts of the MidlandBasin and Northwestern Shelf).

Northwestern Shelf andMatador Uplift

The Woodford Formation is present at mostlocalities on the Northwestern Shelf but is absenton and north of the Matador Uplift (fig. la;pls. 1, 2, 7) (Ellison, 1950; Wright, 1979; Duttonand others, 1982; Ruppel, 1985). In northern LeaCounty, New Mexico, elevation of the Woodfordincreases northward, but the pattern is brokenby several faults (pl. 1). These faults trend north-south or northwest-southeast, generally parallel

low-lying land masses in the latest Devonian orearliest Mississippian and an episode of minorepeirogenic uplift, slight sea-level fluctuations,and brief interruption of marine sedimentation(Stevenson, 1945; Ham and Wilson, 1967; Frezonand Jordan, 1979; Mapel and others, 1979).

to the Central Basin Platform and the axis of theDelaware Basin.

The Woodford thins northwestward acrossEddy County, New Mexico, away from theDelaware Basin (pls. 2, 6), the gradual thinningcoinciding with the increase in elevation (pl. 1).In eastern Chaves and northern Eddy Counties,New Mexico, thin and thick areas are irregularlydistributed and are not clearly related to struc-ture (pl. 2). In the northernmost part of the map,Woodford thickness appears to be structurallycontrolled because isopach contours are orientedeast-west, parallel to the structural trend of theMatador Uplift (pl. 2).

Eastern ShelfThe Woodford Formation was previously

thought not to extend onto the Eastern Shelf(Ellison, 1950; Wright, 1979), but in the presentstudy no clearly defined eastern limit for theWoodford was found (pls. 1, 2). The formationis absent in northeastern Crockett County, inmost of western Irion County, and in a largearea that includes parts of Scurry, Borden, andGarza Counties. However, the formation ispresent across Sterling, Mitchell, and most ofScurry Counties.

The wide spacing of structural contours inthe eastern part of the map (pl. 1) documents agradual increase in elevation of the Woodfordfrom the Midland Basin onto the Eastern Shelf.The Woodford also thins gradually in the samedirection (pls. 2 through 5). On the Eastern Shelf,the Woodford is thin, and the distribution issomewhat irregular and patchy (note thicknessesin Scurry, Mitchell, and Sterling Counties, pl. 2).These structural and isopach trends are uninter-rupted except in southern Irion and northernCrockett Counties, where large-scale faults cutthe section.

10

Page 15: Stratigraphic Analysis of the - University of Texas at Austin

Central Basin Platform andPecos Arch

The Central Basin Platform and Pecos Archare the diverging structural highs in the centerof the map that meet in Crane and northeast-ern Pecos Counties (fig. la; pls. 1, 2). Faultsbounding the Pecos Arch trend east-west,whereas those along the Central Basin Platformtrend northwest-southeast. Some of the larg-est faults, those having throws of a few thou-sand feet, are normal or high-angle reversefaults, although some show evidence of strike-slip motion (Galley, 1958; Walper, 1977;Muehlberger, 1980; Hills, 1984).

The Woodford is absent from the Pecos Archand from many of the faulted structures on theCentral Basin Platform (Ellison, 1950; Galley,1958) (pls. 1, 2, 7). Elevations of the Woodfordor the unconformity representing the Woodfordrange from 980 ft below sea level in northernPecos County to more than 7,000 ft below sealevel in eastern Winkler County. The Woodfordthins over the Central Basin Platform in mostplaces, but in some areas, such as southern Ectorand Winkler Counties, the thickness steadilyincreases westward (pl. 2).

Delaware BasinThe Woodford Formation reaches its maxi-

mum thickness of 661 ft in the Delaware Basinstructural low in western Winkler County(pls. 1, 2). The Woodford is more than 600 ftthick in central and southwestern Winkler,southeastern Loving, and northern Ward Coun-ties. The top of the deepest Woodford is morethan 16,000 ft below sea level in eastern LovingCounty and more than 15,000 ft below sea levelin east-central Reeves County (pl. 1). Severalisolated thick areas whose distribution is faultcontrolled appear in Reeves County (pl. 2); inthe north-central part of the county, thicknesslocally exceeds 500 ft, and in central andsoutheastern areas, 400 ft.

Midland BasinThe axis of the Midland Basin is approx-

imately outlined by the closed -9,000-ft structuralcontours east of the Central Basin Platform in

Texas (pl. 1). The deepest Woodford is nearly9,800 ft below sea level in northeastern GainesCounty (pl. 1). Within the basin thickness trendsare subtle (pl. 2); the Woodford at its thickest is135 ft in north-central Martin County. Two thickareas are indicated by the closed 100-ft isopachcontours in Dawson, Gaines, Andrews, andMartin Counties. Between the thick areas lies aneast-west trend of relatively thin Woodford (50to 100 ft). Another narrow thin trend (<50 ft)lies in southern Martin and southeasternAndrews Counties. These trends parallel struc-tural and isopach trends along the MatadorUplift and Pecos Arch and are at a high angle tothose along the Central Basin Platform imme-diately to the west.

Val Verde BasinThe Woodford is present in the Val Verde

Basin of southern and southeastern Pecos,southern Crockett, northern Terrell, and northernVal Verde Counties (fig. la; pls. 1, 2, 7). Innorthern Brewster and southern Pecos Counties,the Woodford Formation, along with a carbonatesequence typical of the Paleozoic section of thecraton, lies beneath allochthonous rocks of theOuachita Fold Belt (pls. 1, 2, 6 [D–D ,́ well 13]).

Two structural trends are present in the ValVerde Basin (pl. 1). In south-central PecosCounty, faults and structural contours trendnorthwest-southeast, and elevation of theWoodford Formation increases from centralPecos County south westward and westward. InTerrell, southern Crockett, Val Verde, and east-ern Pecos Counties, faults trend east-west, andelevation of the Woodford increases northwardfrom the Ouachita Fold Belt to the Pecos Arch.

In central Pecos County, the Woodford wasinferred to be more than 21,000 ft below sealevel (pl. 1) on the basis of the elevation of theEllenburger Formation (Ewing, 1991), and it wasinferred to be more than 300 ft thick (pl. 2) onthe basis of nearby thickness trends. In centralTerrell County, the Woodford was inferred tobe more than 20,000 ft below sea level, and evendeeper burial is likely beneath the Ouachitaoverthrust (pl. 1). Thicknesses in this part of thebasin locally exceed 400 ft and may be 300 ft ormore in central Terrell County beneath theOuachita allochthon (pl. 2).

11

Page 16: Stratigraphic Analysis of the - University of Texas at Austin

Diablo Platform and WesternOutcrop Belt

The Diablo Platform (fig. la, b) is a majorstructural boundary between the Permian Basinto the northeast and the Chihuahua Tectonic Beltto the southwest that has been strongly affectedby Laramide deformation (Muehlberger, 1980).Most of the faults along the platform trendnorthwest-southeast and follow Proterozoicbasement faults (Walper, 1977; Muehlberger,1980).

LithofaciesTwo lithofacies were identified in the

Woodford Formation—black shale and siltstone.Black shale is pyritic and has parallel laminae;siltstone, a hybrid rock composed of silt-sizedquartz and dolomite grains, is medium to darkgray and has discontinuous and disturbedbedding. Distinguishing between lithologies canbe difficult because differences in color may beslight, and many layers contain subequal mix-tures of silt- and clay-sized grains. Contactsbetween lithofacies may be sharp, particularlyat the base of the siltstones, or gradational; andlithologies are commonly interbedded andinterlaminated.

Lithofacies were defined and described fromcores because weathering had severely oxidizedthe pyrite and organic matter in outcrop. Out-crops are medium to light shades of gray orbrown, whereas cores are black (black shale) orlight to dark gray (siltstone). Textures werefound to be comparable in outcrop and core,and the mineralogy of the silicate and carbonatefraction was similar. Hence, lithofacies analysiswas possible at all localities.

Black ShaleCharacteristic Features

Parallel laminae are the most characteristicfeature of black shale (fig. 4a, b). Other dis-tinguishing features include abundant pyrite, finegrain size, black color, and high radioactivity. Theblack color is caused by high concentrations ofpyrite (as much as 13 vol %; app. C) and organic

The Woodford is absent in the southeastern partof the Diablo Platform, southwest of the map area(Wright, 1979), but it is present in the northeasternpart (Rosado, 1970) (pls. 1,2). The highest observedsubsurface elevation (128 ft below sea level) is innorthwestern Culberson County (pl. 1). The highestoverall elevation (>5,000 ft above sea level) is inthe western outcrops. From the Delaware Basin,elevation of the Woodford gradually increaseswestward across Reeves and Culberson Countiestoward the Diablo Platform (pl. 1), and the forma-tion gradually thins in the same direction (pl. 2).

carbon (as much as 35 vol % [app. C] or 12 wt %TOC; app. D), and high radioactivity is caused byuranium bound in the organic matter (Swanson,1960, 1961; Leventhal, 1981).

Bedding and Sedimentary StructuresContinuous parallel laminae predominate

(fig. 4a), but other stratification types includediscontinuous, wavy, and lenticular laminae andthin beds. Most laminae have no internal struc-ture but can be distinguished by subtle dif-ferences in color that result from differences incomposition (for example, unequal amounts ofdetrital quartz, clay, pyrite, dolomite, andorganic matter and different numbers of sporesand radiolarians). These laminae typically havea varvelike appearance in slabs and thin sections(fig. 4a). Thin graded siltstone-shale coupletswere found in some intervals, mostly in shelfregions (app. B; C3, C4, C13). Most gradedcouplets have sharp bases, and some exhibitprimary sedimentary structures such as fadingripple forms.

Burrows are scarce but commonly causedisrupted or distorted layers (fig. 4a through c).Most burrows are confined to, or start in, silt-stone laminae (fig. 4a, b), but a few were foundexclusively in shale (fig. 4c). Flattened horizontalburrows were the most common type observed,and vertical burrows (fig. 4b, c) were found onlylocally. Burrows are filled by silt, secondarysilica, carbonate, and pyrite in varying propor-tions, and some contain scattered remnants ofanhydrite (fig. 4d).

12

Page 17: Stratigraphic Analysis of the - University of Texas at Austin

Syneresis cracks (fig. 4e) were found locallyon the Central Basin Platform in organic-rich,pyritic black shale. They are short, wide verticalfractures, linear in plan view and wedge shapedin cross section. Syneresis cracks are found inthe middle of the black shale section and not atlithologic or formation contacts (app. B; C2).They are highly compacted and thus are inferredto be syndepositional or very early diagenetic.The cracks are filled mostly by secondary silica(including quartz, chert, and chalcedony) andlocally contain carbonate, along with patchyremnants of anhydrite. Filling must haveoccurred shortly after the cracks formed becausecementing phases are deformed, and the sur-rounding black shale is differentially compacted(fig. 4e). Subaerial exposure is not indicatedbecause pyrite and organic matter in the hostshale are unoxidized.

TextureBlack shale consists of more than 50 percent

clay-sized material and less than 50 percent silt-sized particles (fig. 4a through c). Silt-sizedgrains may be randomly scattered (fig. 4c) orconcentrated in laminae (fig. 4a, b). Lightercolored laminae will typically contain greaterproportions of silt-sized particles than will thedarker laminae.

Median grain sizes of the silt fraction arebetween 0.01 and 0.05 mm. Sand-sized grainsare rare. A few large (as much as 5 cm long)greenish shale clasts exhibiting parallel laminae(fig. 4f) were found locally on the Central BasinPlatform.

CompositionClay-sized material consists of organic matter

and illite, and the silt-sized fraction consists ofmostly dolomite, quartz, pyrite, mica, feldspar,glauconite, biogenic pellets, spores, and radio-larians. Other types of fossils, including con-odonts, brachiopods, trilobites, sponge spicules,and vertebrate debris, were found locally, butonly rarely.

Organic carbon content in core samples rangesfrom 1.4 to 11.6 weight percent TOC (mean =4.5 ± 2.6 wt % TOC for 72 samples), or fromroughly 4 to 35 volume percent organic matter.Outcrops contain much less organic carbon thando the cores (<0.1 to 2.3 wt % TOC; mean =

0.8 ± 0.6 wt % TOC for 25 samples) primarilybecause of oxidation during weathering. Aver-age TOC concentration in each core ranges from2.2 to 9.0 weight percent and in each outcropfrom 0.1 to 1.1 weight percent (app. D).

Organic matter most commonly appears asfine-grained, disseminated, amorphous material(app. D), an oil-generating type. Woody particleswere rare in thin sections and in separatedkerogens. Large plant fragments appeared on afew bedding surfaces in cores from the North-western and Eastern Shelves. Recycled vitriniteoccurs only in black shale from the Central BasinPlatform, Eastern Shelf, and southern MidlandBasin (app. D). The mean reflectance values ofprimary vitrinite in cores and outcrops rangefrom 0.54 to 1.92 percent R (app. D) and repre-sent hydrocarbon generation stages betweenearly oil generation and late wet-gas generation(Hunt, 1979).

Illite is abundant in the black shale. Volumepercentages range between 34 and 79 percentand average 59 ± 3 percent (app. C). The coarseclay mineral fraction (1 to 2 µ m) is detrital illite,whereas the fine clay mineral fraction (<0.2 µ m)is diagenetic illite (Morton, 1985).

Dolomite and quartz are the most commonsilt-sized components. They occur randomlymixed in subequal proportions, and they havethe same grain-size distribution. Dolomite grainsin shale have no overgrowths—most are sub-hedral to anhedral and appear to be abraded(fig. 4g). Euhedral dolomite grains are abundantonly locally in cores from the Central BasinPlatform, Midland Basin, and NorthwesternShelf. Scattered silt- to sand-sized poikilotopicdolomite patches cement clay- or fine silt-sizedparticles in some samples.

Pyrite is ubiquitous in cores and can be foundin a variety of forms: (1) large (as much as 8 cm)nodules (some possessing cone-in-cone fabric),(2) irregular elongate patches, (3) thin streaks,(4) smooth elliptical masses having stromatolite-like or oncolitelike fabric, (5) scattered finegrains, (6) framboids, (7) aggregates (silt sizedor finer), (8) fillings or replacements of minuteorganisms (for example, spores and radio-larians), (9) cement or replacement in burrows,and rarely, (10) fracture fillings. Weathering hasaltered pyrite in outcrop to various oxides andsulfates. Locally, gypsum lines joints and bed-

13

Page 18: Stratigraphic Analysis of the - University of Texas at Austin
Page 19: Stratigraphic Analysis of the - University of Texas at Austin

15

FIGURE 4. Photos of Woodford black shale. (a) Black shale exhibiting continuous parallel laminae at 11,555 ftin the No. 1918 Parks (app. B; Cl, sample C1-10). Note cyclic change from black shale at base, siltstonelaminae increasing upward, and abrupt return to black shale at top. Silty lamina at base, B, is burrowed. TOC =4.2 wt %. (b) Disrupted parallel siltstone laminae in black shale at 10,914 ft in the No. 1 Champeau Federal(app. B; C4, sample C4-4). Disrupted areas, B, are burrows. TOC = 5.0 wt %; Ro = 1.03%. (c) Black shaleexhibiting parallel laminae, scattered silt, and burrows at 12,228 ft in the No. 5 Pacific Royalty (app. B; C9,sample C9-2). TOC = 3.2 wt %. (d) Enlarged view of burrow in photo (c) (see arrow in photo [c]). Burrows arefilled mostly by quartz, Q; pyrite, PY; and patchy remnants of anhydrite, arrows marked A. Rectangular habitindicates quartz is pseudomorph after anhydrite. Crossed nicols. (e) Syneresis cracks filled mostly by silica andcarbonate and locally containing anhydrite at 7,179 ft in No. 43 Yarborough & Alien (app. B; C2, sample C2-7).Note differential compaction of black shale laminae and compactional deformation of syneresis structures.TOC = 8.5 wt % in host shale. (f) Shale clast in black shale at 7,177 ft in No. 43 Yarborough & Alien (app. B; C2,sample C2-6). Clast has parallel laminae, and the outer edge was pyritized, PY. Oblique calcite veins andptygmatic veinlets, V, reflect shearing. TOC = 11.2 wt % in host shale. (g) Silty shale at 7,172 ft (app. B; C2,sample C2-3). Silt is exclusively dolomite, D, and most grains are angular, broken, or abraded. Pellets, P, areelongate fine-grained aggregates and probably biogenic. Plane-polarized light. TOC = 8.5 wt %. (h) Pelletscontaining silt particles at 7,404 ft in the No. 1 Sealy Smith (app. B; C12, sample C12-4). Plane-polarized light.Dolomite/quartz ratio = 1.3/1.0; TOC = 10.2 wt % ; Ro = 0.55%. (i) Pellets composed mostly of clay at 11,639 ftin the No. 1 Walker (app. B; C5, sample C5-2). Silt grains are dolomite and quartz. Plane-polarized light.Dolomite/quartz ratio = 1.3/1.0; TOC = 2.8 wt %. (j) Pellets and flattened spores at arrows from same thinsection as those in photo (h). Spores are Tasmanites. Plane-polarized light.

Page 20: Stratigraphic Analysis of the - University of Texas at Austin

ding planes, and iron oxide appears as pyritepseudomorphs, indicating that these rocks werehighly pyritic before weathering. Elsewhere,disseminated ferric oxides record the formerabundance of pyrite in the Percha and Sly GapFormations.

Muscovite flakes appear in all samples. Micaflakes and illite typically are well orientedparallel to bedding. In biogenic pellets, however,illite may comprise domains of differing orien-tations. Locally some mica flakes lie at highangles to bedding, a few flakes being oriented90 degrees to bedding. Such flakes appear to bepart of larger clumps of organic-bound sediment.

Feldspar (microcline) and glauconite are rarein black shale. Feldspar appears mostly insamples from the Northwestern Shelf, CentralBasin Platform, and western Midland Basin(app. C; Cl, C2, C4, C13). Glauconite occurs asan isolated grain or two in many thin sections.

Biogenic pellets are common (as much as 11%;app. C) in many black shale samples. Theyappear as flattened silt sized to fine sand sizedaggregates and impart a microlenticular fabricto the rock when viewed in thin section (fig. 4hthrough j). Pellets are easily distinguished fromburrows in plan view because pellets exhibit notrail-like patterns on bedding surfaces or in crosssection because pellets show no cross-cuttingcontacts or internal stratification. Most pelletsconsist of illite, but some consist of silt-sizedgrains of quartz and dolomite (fig. 4h). Silty pel-lets commonly are cemented by carbonate andare flattened slightly less than clay pellets.

Spores are minor components in black shale,but they are widely distributed. Generally sporesare flattened as a result of compaction (fig. 4j).However, in some intervals spores have beenreplaced by pyrite or infilled by pyrite, chert, orcarbonate. Locally, early infilling is indicated byspores that are uncompacted or only slightlyflattened.

Radiolarians also are a minor but widelydistributed component. They are composedmostly of chert or chalcedony, but some havebeen partly or completely replaced by pyrite orcarbonate. Spores and radiolarians may berandomly scattered throughout a laminated se-quence or concentrated in laminae or thin beds.Radiolarian chert layers were observed locally

on the Central Basin Platform and in thesouthern Midland Basin (app. B; C2, C6).

Trilobite fragments are sparsely scattered insome intervals and locally occur alongsidepellets. Most are carbonate, but a few have beenpartly replaced by chert. Trilobite fragmentslocally are common at the top of the formationalong the disconformity with the Mississippianlimestone (app. B; C10).

Brachiopods are scarce in the Woodford.Inarticulate brachiopods (Lingula) were recog-nized on bedding surfaces in cores from theCentral Basin Platform and the NorthwesternShelf (app. B; C2, C4, C9). One silicified articulatebrachiopod was found in black shale on theCentral Basin Platform (app. B; C2). Elsewhere,articulate brachiopods are abundant only locallyat the top of the formation (app. B; C10).

Phosphatic fossil debris is a minor componentin black shale. Conodonts are scarce but widelydistributed, and bone and teeth fragments andfish scales also are rare. Phosphatic debrissometimes is concentrated in the siltier shalesand in interstratified siltstones.

Sponge spicules were found only locally inone core from the Central Basin Platform wheremonaxons were scattered parallel to stratification(app. B; C2). All of the spicules had altered tochert.

Secondary silica is the major constituent insome layers and sedimentary structures. Sec-ondary silica in the form of chert, chalcedony,and megaquartz fills or replaces fossils andcements or replaces burrows and syneresiscracks (fig. 4c through e). Megaquartz that haspseudomorphic rectangular cleavage afteranhydrite was found locally associated withanhydrite (fig. 4d). Also, some of the chalcedonyin burrows and syneresis cracks is length-slow,suggesting that it replaced evaporites (Folk andPittman, 1971).

SiltstoneCharacteristic Features

Siltstone in the Woodford Formation is ahybrid siliciclastic-carbonate rock in whichdolomite and quartz are the dominant silt-sizedframework grains. Compared with black shale,siltstone has coarser grain size, lighter color,

16

Page 21: Stratigraphic Analysis of the - University of Texas at Austin

more disrupted or discontinuous strata, andlower radioactivity. Siltstone, unlike the car-bonate lithologies of bounding formations, hasa more uniform silt-sized texture, abundantquartz grains, no chert, no large body fossils,and higher radioactivity.

Bedding and Sedimentary StructuresStratification ranges from thin laminae to

thin beds. Continuous, discontinuous, and wavyparallel laminae commonly are preserved, butstratification typically is disrupted by burrowing(fig. 5a, b) or, more rarely, contorted by soft-sediment deformation (fig. 5b, c).

Interbedded and interlaminated dark-gray toblack shale, fine-grained dolomite grainstone,fine-grained lime grainstone, and lime mudstonelocally are common (app. B; C9, C11). Theinterbedded shales and mudstones typicallyexhibit continuous, discontinuous, or wavyparallel laminae, and the grainstones, discon-tinuous and disturbed layers.

Most siltstones and grainstones have sharplower contacts (fig. 5d through f), and many formgraded couplets with shale (fig. 5e, f). Othershave gradational upper and lower contacts(fig. 5g). Cores containing well-developed silt-stone lithofacies commonly consist of verticallystacked couplets in which siltstone beds as thickas 10 to 15 cm grade upward into shale layersas thick as 5 cm (fig. 5a, b, e). Primary sedi-mentary structures include normal grading(fig. 5a, b, e, f), fading ripple forms (fig. 5e),climbing ripple cross-stratification (fig. 5d),horizontal stratification, soft-sediment defor-mation (fig. 5b), and flow-sheared laminae(fig. 5c). The vertical succession of structurestypically comprises a partial or complete Boumasequence (fig. 5e). Siltstone sequences such asthese constitute a basal facies of the Woodfordin the Northwestern Shelf and northern MidlandBasin (app. B; C5, C9, C11).

TextureMedian grain sizes of siltstone are between

0.01 and 0.05 mm. Typically, little or no sand-sized material is present, although sand grainsas. large as the medium-sized grade wereencountered locally. Clay-sized material rangesfrom 0 to almost 50 percent by volume. Silt-

stone is moderately to poorly sorted and rarelywell sorted.

CompositionQuartz and dolomite are the most abundant

framework grains in siltstone, and they typicallyhave the same grain-size distribution (fig. 5h).They are commonly present in subequal pro-portions and are mixed with a variety of othercomponents so that neither constitutes more than50 percent of the rock. Dolomite is mostly sub-hedral or anhedral, and such grains commonlyhave an abraded appearance (fig. 5h). Euhedralgrains were found locally, but many haveanhedral or subhedral cores rimmed by euhedralovergrowths. In most siltstones, dolomite formsan interlocking mosaic with quartz, yet dolomiteis rarely poikilotopic, even in dolomite grain-stones. Locally, poikilotopic patches of dolomitecement a few angular silt-sized grains.

Other silt-sized constituents are pyrite, mica,feldspar, glauconite, phosphatic debris, and rarezircon and tourmaline. Pyrite is common andappears as nodules, euhedral crystals, irregulargrains, aggregates, and framboids. In some beds,pyrite has subhedral and anhedral shapes similarto those of quartz and dolomite (fig. 5f, h) andmay be reworked.

Mica (muscovite) was observed in all quartz-dominated siltstones and most dolomite-dominated siltstones; however, it is rare incarbonate mudstones and grainstones. Feldspar(microcline) is a minor component mostly inquartz-dominated siltstone. Both mica andfeldspar are more abundant in the NorthwesternShelf and northern Midland Basin (app. B andC; C5, C11) where they occur along with minoramounts of the ultrastable silicates zircon andtourmaline.

One or two grains of glauconite were seen inmany samples, but glauconite is concentratedonly locally at the top of the formation (app. B;C10, C13). Many core samples contained minoramounts of phosphatic debris, mostly conodontsand fish debris.

Illite and organic matter compose the finefraction of siltstones, and in some samples theclay constitutes almost 50 percent of the rock.Illite is abundant in wispy laminae, in the upperpart of graded layers, and in gradational shaly

17

Page 22: Stratigraphic Analysis of the - University of Texas at Austin
Page 23: Stratigraphic Analysis of the - University of Texas at Austin

19

FIGURE 5. Photos of Woodford siltstone. (a) Burrowed quartz-dominated siltstones and interlaminated shalein No. 1 Walker at 11,681 ft (app. B; C5, sample C5-12). Mean TOC of interlaminated interval = 0.2 wt %.(b) Quartz-dominated siltstone in No. 1 Williamson at 13,064 ft (app. B; C11, sample C11-10). Note soft-sediment deformation fabric just below pyrite, PY, nodules. Mean TOC of interlaminated interval = 0.7 wt %.(c) Core chip showing fine-grained dolomite grainstone in No. 1 Federal Elliott at 14,638 ft (app. B; C13,sample C13-6). Contorted laminae record flow shear during rapid deposition in a bottom flow. Dolomite/quartz ratio = 40/1. (d) Very thin dolomite-dominated siltstone bed in No. 1 A. E. State at 13,771 ft (app. B; C3,sample C3-6). Bed contains small-scale climbing ripple cross-laminae and grades into silty shale at top. Darkpatches, PY, are pyrite. TOC in underlying shale bed = 2.3 wt %. (e) Dolomite-dominated siltstone laminaein No. 1 A. E. State at 13,768 ft (app. B; C3, sample C3-5). Middle lamina shows Bouma sequence that hasgraded, A; flat, B; and rippled, C, intervals. Ripple crests are spaced roughly 1.5 cm apart. Mean TOC in shalelaminae = 1.6 wt %. (f) Enlarged view of graded interval, A, in photo (e). Silt is a subequal mixture of dolomite,quartz, and pyrite. (g) Dolomite-dominated siltstone laminae in No. 43 Yarborough and Alien at 7,172 ft(app. B; C2, sample C2-3) having indistinct contacts and lacking internal structure. (h) Magnified view of quartz-dominated siltstone shown in photo (b). White and dark-gray grains are quartz, pale-gray grains are dolomite,black grains are pyrite. Note angular and abraded appearance of some dolomite grains. Dolomite/quartzratio = 0.95/1.0. Crossed nicols.

Page 24: Stratigraphic Analysis of the - University of Texas at Austin

intervals between black shale and siltstone(fig. 5a, b, d through g).

Organic matter is not abundant, and siltstonecores and outcrops average less than 1 weightpercent TOC (app. D). In individual samplesTOC concentrations range between 0.1 and1.1 weight percent (mean = 0.5 ± 0.3 wt % TOCfor 20 samples), which roughly corresponds to0.3 to 3 percent organic matter by volume. Thetypes of organic matter include amorphousparticulate material, spores, and wood, butamorphous organic matter greatly predominatesin all samples. Spores are rare, and only a fewwood fragments were found on bedding planes.Siltstones contain only small amounts of primaryvitrinite and no recycled vitrinite (app. D),suggesting that terrigenous source areas hadminimal plant cover and few carbonaceousrock exposures. Reflectance values range from0.8 percent to 1.3 percent and are directly relatedto present-day burial depth (app. D).

Formation-Boundary LithologiesLower Contact

In the Permian Basin, contact betweenSilurian-Devonian carbonate rocks and theoverlying Woodford Formation was preservedin two cores, the No. 1 A. E. State and the No. 1Walker (app. B; C3, C5). In the No. 1 Walkercore, the Woodford disconformably overliesSilurian-Devonian limestone that consists ofmottled fine-grained grainstones and brachiopodgrainstones (fig. 6a, b) that contain scatteredchert lenses and nodules. The upper surface ofthe limestone is irregular, and locally it is bored.The basal Woodford layer is conglomeratic chertarenite that contains glauconite and phosphaticdebris (fig. 6a, b) and is texturally and com-positionally similar to the basal chert breccia inthe Arbuckle Mountains described by Amsden(1975, 1980). Phosphatic debris includes cono-donts, assorted fragments (bone, teeth, fishscales, Lingula), aggregates (fecal material), andooids that exhibit both radial-fibrous and con-centric fabric. Basal Woodford chert arenite isunsorted and has no current-induced primarysedimentary structures; thus it appears to be aresidual lag produced by dissolution of theunderlying cherty limestone. The fossils, glau-

conite, and phosphatic ooids indicate open-marine conditions and slow sedimentation.

In the No. 1 A. E. State core (from Lea County,New Mexico) brecciated, cavernous limestone isoverlain disconformably by black shale. Theuppermost 1 ft of limestone contains black shaleclasts, and the basal Woodford contains scatteredangular fragments of black shale and limestone(app. B; C3). The transition from limestone toblack shale is abrupt; however, the contact isirregular and penetrative, and infiltration ofmud tens of feet downward into the underlyinglimestone has occurred. Some of the solutioncavities and fissures in the limestone are partlyor completely filled by black shale that eitherhas no structure or contains deformed, contortedlaminae indicative of soft-sediment deformation(fig. 6c). The shale-filled cavities and fissures atLea County, New Mexico, are similar to thosein Andrews and Terry Counties, Texas, de-scribed by Peirce (1962).

Upper ContactContact between the Woodford and the over-

lying Mississippian limestone was preserved intwo cores, the No. 1 Brennand and Price andthe No. 1 Federal Elliott (app. B; C10, C13). Inthe No. 1 Brennand and Price, the uppermostWoodford contains articulate brachiopods,trilobite fragments, black shale clasts, dolo-mite grains, glauconite, and phosphatic debris(fig. 6d). The overlying Mississippian limestonesare mostly laminated fine-grained grainstonesalong with some lime mudstones, sparselyfossiliferous grainstones, wackestones, andpackstones. Locally these carbonate lithologiescompose thin, graded beds. Chert beds, lenses,and nodules, locally spiculitic, are scatteredthroughout the Mississippian limestone section.Contact between the black shale and the over-lying carbonate rocks is sharp and discon-formable (fig. 6d), marking an abrupt change inlithology and fauna.

In the No. 1 Federal Elliott, Mississippianlimestone rests conformably on the Woodford.Woodford black shale grades upward through10 ft of interbedded dark-gray lime mudstone,black siltstone, and black glauconitic sandstoneinto medium to dark-gray fine-grained Missis-sippian grainstones and lime mudstones (app. B;

20

Page 25: Stratigraphic Analysis of the - University of Texas at Austin

21

FIGURE 6. Photos of Woodford contacts, (a) Core slab showing lower contact, arrow, in No. 1 Walker at11,689 ft (app. B; C5, sample C5-14). Silurian-Devonian limestone, DLS, overlain by basal Woodford chertarenite, CA, and black shale, BS, at top of core, (b) Thin-section photomicrograph of area at arrow in (a). Chert,C; pyrite, PY; phosphate, P. Three phosphatic grains from left to right are ooid containing a chert nucleus,aggregate of probable fecal origin, and skeletal fragment. Finer grains are chert (light) and phosphate (black).Below contact is fine-grained grainstone. Crossed nicols. (c) Core chip from 13,850 ft in No. 1 A. E. State(app. B; C3). Woodford black shale, W, in solution cavities in Silurian-Devonian limestone. Contorted laminaein shale and shale penetrating limestone crevices at arrows indicate soft-sediment infiltration of mud into theunderlying limestone, (d) Thin-section photomicrograph of Woodford-Mississippian contact, arrow, at 8,459 ftin No. 1 Brennand and Price (app. B; C10, sample C10-1). Upper Woodford consists of brachiopod shells andtrilobite carapaces (white ribbonlike material), black shale clasts, and silt-sized grains (white specks) that aremostly dolomite. Fine white streaks in black shale at base are brachiopod and trilobite fragments. Mississippianabove contact is fine-grained grainstone, G, and chert bed containing scattered, unreplaced remnants of carbonate,C. Crossed nicols.

Page 26: Stratigraphic Analysis of the - University of Texas at Austin

C13). This 10-ft interval was assigned to theWoodford Formation because it is markedlymore radioactive than the overlying rocks andbecause it contains diagnostic Woodford featuressuch as varvelike parallel laminae and abundantpyrite.

High concentrations of glauconite and phos-phate in sedimentary rocks indicate low sedi-mentation rates (Odin and Letolle, 1980). Thetop stratum of the Woodford at these twolocalities is consequently inferred to haveaccumulated more slowly than the rest of theformation. Commonly, glauconitic grains andphosphatic ooids and pellets are unbroken,current-induced primary sedimentary structuresare absent, and brachiopods possess articulatedvalves, indicating little or no active sedimenttransport at the close of Woodford deposition.The abundance of reduced iron, sulfur, and car-bon and the absence of oxidized phases docu-ment absence of oxidation and imply absence ofsubaerial exposure. The upper boundary at thesetwo localities thus suggests a submarine hiatusduring which sedimentation slowed or ceasedbut the sea floor did not emerge.

Lithofacies CorrelationBasal siltstone in Woodford cores from the

Northwestern Shelf and northern Midland Basin(app. B; C5, C9, C11) is herein correlated withthe lower Woodford unit of Ellison (1950) onthe basis of lithology, radioactivity, and strati-graphic position (fig. 7). Basal siltstone, whichis a hybrid of silt-sized quartz and dolomite, iscomparable to Ellison’s lower unit in its highcarbonate content and low radioactivity. Unfor-tunately, Ellison’s cores were discarded, anddirect comparison of lithologies was impossible.In the subsurface, both the lower unit and thebasal siltstone immediately overlie the regionalunconformity.

Stratigraphic position and lithology also sug-gest correlation of basal Woodford siltstone withthe upper Middle to lower Upper DevonianOnate Formation in southeastern New Mexico.Both units rest on the regional unconformitysurface and comprise a stratigraphic successionof interbedded siltstone, carbonate, and shale inwhich dolomitic siltstone is the dominant

lithology. The proposed correlation is consistentwith that by Wright (1979), who suggested acorrelation of the lower unit of Ellison (1950)with the Ives Breccia Member of the upperMiddle Devonian to Lower Mississippian HouyFormation in Central Texas.

Basal siltstone also occupies the same strati-graphic position above the regional uncon-formity as the Canutillo Formation in West Texasand the Misener and Sylamore Sandstones inOklahoma and Arkansas. These formally namedunits are mostly late Middle to early LateDevonian in age and are locally as young asEarly Mississippian. Although the units arediachronous across the southern Midcontinent(Amsden and others, 1967; Freeman andSchumacher, 1969; Rosado, 1970; Amsden andKlapper, 1972; Amsden, 1975), they are at leastpartly correlative.

The black shale lithofacies is correlated withthe middle and upper Woodford units of Ellison(1950) also on the basis of lithology, radioactivity,and stratigraphic position (fig. 7). Ellison’smiddle and upper units are not described asseparate lithofacies in this report because striking,lithologic differences between them are absentin cores (app. B; C1, C6). Although both unitsare pyritic black shale exhibiting parallellaminae, the middle unit is more radioactive(Ellison, 1950); hence, the middle and upperunits can be mapped using gamma-ray logs(fig. 7; pls. 3 through 7). Wright (1979) correlatedthe middle unit with the Doublehorn ShaleMember and the upper unit with the unnamedphosphatic member of the Houy Formation inCentral Texas, thereby implying that the upperunit is partly Kinderhookian. Wright’s correla-tion seems reasonable because the middle unitin both formations has higher radioactivity andmore spores than does the upper unit (Ellison,1950; Cloud and others, 1957).

Well log correlations (pls. 3 through 7) showthat complete Woodford intervals containing allthree units of Ellison (1950) are common only inthe Midland, Delaware, and Val Verde Basins(pl. 3, A–A´, wells 4, 9; pl. 4, B–B´, well 5; pl. 5, C–C´, wells 2 through 6, 9, 10; pl. 6, D–D´, wells 9through 12; pls. 7, E–E´, wells 10, 11, 16, 17).Elsewhere Woodford sections are incompletemostly because of the absence of the lower or

22

Page 27: Stratigraphic Analysis of the - University of Texas at Austin

upper units. The lower unit gradually pinchesout and is overstepped by the middle unit alongthe basin flanks (pls. 3 through 7), indicatingdepositional onlap. Lines of section showingonlap include (1) from the Midland Basin towardthe Eastern Shelf (pl. 3, A–A´, wells 9 through 13;pl. 4, B–B´, wells 11 through 15; pl. 5, C–C´, wells9 through 12), (2) from the Midland Basin ontothe Pecos Arch (pl. 7, E–E ,́ wells 11 through 14),(3) westward from the Delaware Basin towardthe Diablo Platform (pl. 4, B–B ,́ wells 1 through3; pl. 5, C–C ,́ wells 1, 2), and (4) in the westernMidland Basin (pl. 4, B–B´, well 10). Manysections in which the upper unit is absent areoverlain by Mississippian limestone, indicatingnondeposition or erosional truncation thatoccurred after Woodford deposition but beforeMississippian limestone deposition. Sectionsshowing truncation include (1) along the easternmargin of the Central Basin Platform (pl. 3, A–A´, well 8; pl. 4, B–B ,́ well 9), (2) in the northernand central Midland Basin (pl. 7, E–E ,́ wells 7, 8,11), and (3) on the Northwestern Shelf (app. B;C5, C9). Most of the lines of section that showonlap also show evidence of increased truncationof the Woodford in the direction of onlap,suggesting that these were the last flooded andfirst exposed areas during the Late Devonian

transgression and latest Devonian regression.The patterns of onlap and truncation (pls. 3through 7) indicate that all of the structuralprovinces shown in figure la had topographicexpression in the Late Devonian. Onlap in thewestern Midland Basin supports the observationof Galley (1958) that a middle Paleozoic pre-cursor of the Central Basin Platform lay slightlyto the east of the present-day structure.

Lithofacies DistributionCorrelations shown in the cross sections (pls. 3

through 7) and the Woodford lithofacies dis-tribution shown in a fence diagram (fig. 8) revealthat black shale is nearly ubiquitous and themost widely distributed lithofacies. Siltstone ismore common in the northern part of the studyarea and in basinal depocenters. Silt-sized quartzis more abundant in northern and eastern areas,and silt-sized dolomite is more abundant in thefar western outcrop belt and along the CentralBasin Platform. Log correlations indicate thatbasal siltstone is areally restricted to deep partsof the Delaware, Midland, and Val Verde Basins,proximal areas on the Northwestern Shelf, anda few localities on the Central Basin Platform(pl. 3, A–A´, wells 3 through 10; pl. 4, B–B´, wells

23

FIGURE 7. Log correlation of Woodford lithofacies. Reference log from Ellison (1950). Datum is top of Woodford.For detailed core descriptions see appendix B (Cl, C5, C6, C9).

Page 28: Stratigraphic Analysis of the - University of Texas at Austin
Page 29: Stratigraphic Analysis of the - University of Texas at Austin

2 through 5, 7, 9, 11, 12; pl. 5, C–C ,́ wells 2through 6, 9, 10; pl. 6, D–D ,́ wells 6 through 12;pl. 7, E–E ,́ wells 7, 8, 10, 11, 16, 17).

Facies changes between black shale and silt-stone appear in many parts of the study area(fig. 8; pls. 3 through 7). Siltstone beds commonthroughout the Sly Gap Formation in south-eastern New Mexico correlate with black shalein the Percha and Woodford Formations to thesouth and east (Laudon and Bowsher, 1949;Ellison, 1950; Rosado, 1970). Dolomitic siltstonesof the Onate Formation in New Mexico alsocorrelate with dolostone and cherty dolostonebeds of the Canutillo Formation in West Texasand with black shales in the Percha andWoodford Formations (King and others, 1945;Rosado, 1970). On the Northwestern Shelf,siltstone is the basal unit of the Woodford atsome localities (app. B; C5, C9), but it is higherin the section at others (app. B; C3, C13). In theDelaware and Val Verde Basins, siltstone bedsappear to be common throughout the forma-tion, as indicated by the generally reducedradioactivity and the highly erratic nature of thegamma-ray log patterns shown in plates 5through 7 (pl. 5, C–C ,́ logs 4, 6; pl. 6, D–D ,́ logs9 through 13; pl. 7, E–E ,́ logs 16, 17).

Depositional ProcessesSiltstone

Many of the siltstone strata and siltstone-shalecouplets in the Woodford Formation (fig. 5athrough f) closely resemble the silt and mudturbidites described by Piper (1978) and Stow andPiper (1984) and the distal storm deposits describedby Aigner (1982, 1984). In the Woodford, thesestrata range from laminae less than 2 mm thick tobeds rarely more than 10 to 15 cm thick. Theycommonly contain graded layers (fig. 5a, d throughf), climbing ripple cross-stratification (fig. 5d),horizontal stratification, fading (incipient) rippleforms (fig. 5e), flow-sheared laminae (fig. 5c), andlaminae contorted by soft-sediment failure (fig. 5b).Many of these strata are partial or complete Boumasequences that have scoured bases, normallygraded sequences, and a vertical succession ofprimary sedimentary structures that indicate rapiddeposition from a waning current during a singleevent.

Both fine-grained turbidites and distal stormdeposits described in the literature have similarthicknesses and sedimentary structures (Piper,1978; Aigner, 1982, 1984; Stanley, 1983; Stow andPiper, 1984; Schieber, 1987; Davis and others,1989). Mud turbidites in the deep ocean consistof the division E mud of Bouma (1962), whichPiper (1978) subdivided into laminated, graded,and ungraded units. The vertical pattern andthe contained sedimentary structures, such asgrading and low-amplitude climbing ripples, arediagnostic of turbidite origin (Stow and Piper,1984). Silt turbidites are silt-dominated sequencesthat exhibit the same suite of sedimentary struc-tures and the same divisions (Bouma A throughF) as classical sandy turbidites (Stow and Piper,1984). The siltstone and shale layers in theWoodford Formation (fig. 5a through f) differfrom silt and mud turbidites described in theliterature (for example, Piper, 1978; Stanley, 1983;Stow and Piper, 1984) only in the scarcity ofbioturbation in the shale that is common at thetop of the turbidite sequence (Bouma division Emud and division F pelagite). This differenceindicates that anoxic bottom conditions toxic tobenthic organisms prevailed throughout thebasin during deposition of the shale laminae.

Sedimentary processes related to storms, suchas wind-forced currents (Morton, 1981), ebbcurrents produced by storm surge setup (Nelson,1982), and seaward-flowing currents caused bycoastal down welling (Swift and others, 1983),deposit sediment that has textures and struc-tures virtually identical to those of turbidites.Distal storm deposits characteristically are finegrained, thinly stratified, and normally graded,having scoured bases and Bouma sequences(Aigner, 1982). They differ from proximalequivalents in grain size and layer thickness andin their having no hummocky stratification oroscillatory ripples, both of which, when present,indicate deposition under combined flow con-ditions above wave base (Aigner, 1982). Whetherstorms produce turbidity currents is debatable,but it is clear that storms generate bottom cur-rents that transport large quantities of sediment(Hayes, 1967; Morton, 1981, 1988; Nelson, 1982;Walker, 1984, 1985). Storm-generated bottomflows and turbidity currents may represent endmembers of a single process if, as suggested by

25

Page 30: Stratigraphic Analysis of the - University of Texas at Austin

Walker (1984, 1985), distal storm currents pass-ing below wave base become turbidity currents.Such a subtle change in the transport mechanismmay explain the present difficulty in distin-guishing fine-grained turbidites from stormdeposits in the stratigraphic record. Whetherthe siltstones and siltstone-shale couplets inthe Woodford Formation are turbidites orstorm deposits is likewise problematic, but thepresence of grading and partial or completeBouma sequences indicates deposition frombottom flows.

Black ShaleMost layers in the black shale lithofacies

(fig. 4a through h) do not have grading or Boumadivisions as do beds in the siltstone. Blackshale that displays undisturbed parallel laminaetypically contains higher concentrations ofmarine organic matter, less clastic material, andmore planktonic microfossils (for example,radiolarians, spores, conodonts) than do theBouma E and F shales of the siltstone-shalecouplets (fig. 5a through f). Shale displayingparallel laminae constitutes the bulk of theWoodford black shale lithofacies and is mostlypelagic in origin.

Origin of the thin varvelike siltstone and shalelaminae in pelagic black shale (fig. 4a, b, e) isless certain. These laminae may representmud turbidites or storm layers too small orfar from the source to produce grading andrecognizable Bouma divisions, or they mayrepresent episodic fallout from the pycno-cline. Pierce (1976), Maldonado and Stanley(1978), and Stanley (1983) described detachmentof low-concentration turbidity plumes andentrainment of the muddy water along theisopycnals in strongly density stratified watercolumns. Episodic fallout of material (forexample, terrigenous silt and planktonic tests)occurs as particle concentration builds upand exceeds the density of the pycnocline,producing a relatively clean, well-sorted,structureless lamina of widespread arealextent. Similar laminae are common in muddymarine sediments, such as those found in theeastern Mediterranean Sea near the Nile delta(Maldonado and Stanley, 1978; Stanley, 1983).Sediment deposition by this process seems likely

during Woodford accumulation because ofthe exclusively fine grained texture of the rocksand because of the strong density stratificationthat existed within the basin. Water-densitystratification is an inherent property of the seaand, judging from the scarcity of bioturbationand its implicit link with bottom stagnation andanoxia (Byers, 1977; Arthur and Natland, 1979;Demaison and Moore, 1980; Leggett, 1980;Ettenshon and Barron, 1981; Stanley, 1983; Pratt,1984; Ettensohn and Elam, 1985; Stein, 1986;Davis and others, 1989) strong density stratifica-tion probably occurred during Woodford blackshale deposition. (See also Paleoceanography,p. 33.) In this context, the relative abundancesof benthic fossils, trace fossils, and undisturbedparallel laminae in the Woodford (figs. 4athrough h, 5a through h) indicate that the blackshale and siltstone lithofacies represent anaerobicand dysaerobic biofacies, respectively (Rhoadsand Morse, 1971; Byers, 1977).

Lithologic Patterns and Origin ofSediments

The Woodford Formation consists of varyingproportions of terrigenous, pelagic, and authi-genic constituents (app. C), and textural andcompositional evidence indicates much resedi-mentation within the basin. Terrigenous materialincludes fine-grained quartz, muscovite, micro-cline feldspar, illite, wood and leaf fragments,vitrinite, and the trace heavy minerals (zirconand tourmaline). The silt-sized silicate mineralsare most common in the northern basin. Locally,in rocks from the Northwestern Shelf, coarse-grained mica flakes glitter on fresh beddingsurfaces (app. B; C4, C5), and the silt-sizedfraction is subarkosic (app. C). The distributionand texture of these minerals indicate that theprincipal source was the land north of the basin,the Pedernal Massif and northern Concho Arch(fig. 1b).

Siltstone depocenters lie in the northern, central,southern, and westernmost parts of the basin(fig. 9a) in areas coincident with the modern-dayNorthwestern Shelf, the deepest parts of theDelaware, Val Verde, and Midland Basins, andthe Sacramento Mountains. The patchy distributionof these depocenters suggests that sediment

26

Page 31: Stratigraphic Analysis of the - University of Texas at Austin

FIGURE 9. Regional lithologic variations in Upper Devonian rocks in West Texas and southeastern NewMexico. Maps show Late Devonian shoreline and limit of Tobosa Basin depocenter (fig. Ib). (a) Siltstonedepocenters, (b) Illite depocenters, (c) Recycled vitrinite depocenters, (d) Radiolarian chert depocenters,(e) TOC concentration in black shale, (f) Depocenters of silt-sized dolomite where dolomite/quartz ratio isgreater than 1. (g) Depocenters of dysaerobic, shallow-water sedimentary structures.

27

Page 32: Stratigraphic Analysis of the - University of Texas at Austin

bypassing was common as silt and mud movedfrom siliciclastic source areas downslope into thebasin. This inference is consistent with theinterpretation that most silt was deposited frombottom flows, a mechanism sensitive to bottomirregularities and channelization.

The most abundant terrigenous componentin the Woodford is illite (app. C). Detrital illitehas an apparent Rb-Sr source age of 540 m.y.,an age uncommon in North American basementrocks but common in regionally metamorphosedrocks found in large areas of Africa and SouthAmerica (Morton, 1985). The source age and thegood fit to the isochron for data from widelydifferent localities in West and Central Texasare cited as evidence that illite came from asouthern (Gondwana) source or was thoroughlymixed during transport from multiple sources(Morton, 1985). In the present study, the highestconcentrations of illite (>60%) were found innorthern, southeastern, and westernmost regions(fig. 9b) in the present-day Northwestern andEastern Shelves, Midland Basin, and westernoutcrop belt. The wide distribution of illitedepocenters and their proximity to northernsiltstone depocenters and siliciclastic sourceareas suggest that illite came from multiplesources on the Pedernal Massif and Concho Arch(fig. 9b). The broad extent of the exposed landimplies that it derived from diverse stratigraphiclevels. Although contribution from a Gondwanasource cannot be ruled out because of theabsence of control in the southern part of thebasin, a mixed provenance for illite seems mostlikely.

Trace amounts of vitrinite are ubiquitous,documenting a small contribution of land plantdebris to all parts of the basin. Recycled vitrinitewas found only in the eastern and central partsof the basin (fig. 9c) in black shale from theCentral Basin Platform, southern Midland Basin,and Eastern Shelf (app. D), indicating that theseareas were close to emergent land that displayederoding bedrock. A few wood and leaf impres-sions were found mostly in northern and east-ern parts of the basin in rocks from theNorthwestern and Eastern Shelves (app. B; C4,C7, C13). Their distribution implies that landareas on the Pedernal Massif and Concho Arch

supported most of the terrestrial plant life inthe study area during the Late Devonian.Abundances of vitrinite and land plant remainsare low, however, even in the siltstones, indi-cating that terrestrial source areas were onlysparsely vegetated.

Pelagic constituents include radiolarians,amorphous particulate organic matter, algalspores, conodonts, fish fragments, and associatedfecal material. Radiolarian chert is most commonin the central and eastern parts of the basin(fig. 9d) at localities on the present-day CentralBasin Platform and in the southeastern MidlandBasin (app. B; C2, C6). Chert is also abundant inthe Canutillo Formation in West Texas (King andothers, 1945; Rosado, 1970). Anomalously highbiogenic silica is perhaps the best indication ofnutrient-rich water upwelling in ancient seas(Parrish and Barron, 1986; Hein and Parrish,1987) and suggests that upwelling occurred inthe basin and was most pronounced in centraland western areas. Intrabasinal upwelling is alikely consequence of the major oceanic up-welling that occurred adjacent to the study areaalong the margin of the North American cratonduring the Late Devonian. This upwelling epi-sode is recorded as extensive Upper Devoniannovaculite beds of biogenic origin in theOuachita allochthon (Park and Croneis, 1969;Lowe, 1975; Parrish, 1982).

Volumetrically, amorphous organic matter(AOM), which accounts for nearly all of the TOC,is the most abundant pelagic constituent in theWoodford (app. C). The highest TOC concen-trations (>6 wt %) are found in the center ofthe basin on the modern-day Central Basin Plat-form (fig. 9e). Somewhat lower TOC values (4 to6 wt %) are found to the east and north in areascoincident with parts of the western and east-ern Midland Basin, southern NorthwesternShelf, and western margin of the Eastern Shelf(fig. 9e). Localities that have the highest TOCconcentrations also have the most radiolarianchert, suggesting that high TOC values recordincreased biologic productivity at sites of intra-basinal upwelling. The area that has the highestTOC’s (fig. 9e) is surrounded by siltstone depo-centers (fig. 9a), supporting the inference that itwas bypassed by siliciclastic sediment.

28

Page 33: Stratigraphic Analysis of the - University of Texas at Austin

Authigenic material includes dolomite, pyrite,secondary silica, glauconite, anhydrite, calcite,and phosphatic ooids. Some cored intervals onthe Central Basin Platform and NorthwesternShelf (app. B; C2, C3) contain abundant pristine,euhedral dolomite rhombs floating in organic-rich black shale. The texture and association aresimilar to those observed in Deep Sea DrillingProject (DSDP) cores and in very young sedi-ments in the Gulf of California (Baker andKastner, 1981), suggesting that the rhombs areauthigenic and formed in situ. Most of thedolomite in the Woodford, however, appears tobe resedimented because it contains abradedanhedral and subhedral silt-sized grains andcommonly appears randomly mixed withquartz in graded layers and Bouma sequences(fig. 5e, f, h). Derivation from ancient dolomiticrocks is not indicated. The poor durability ofdolomite precludes long-distance subaerialtransportation. Moreover, dolomite in theWoodford is typically monocrystalline andmonotonously uniform in texture, whereas inolder Paleozoic rocks, dolomite texture is quitevariable. One would expect to see dolomitic rockfragments and a greater variety of textures ifWoodford dolomite were terrigenous detritus.

If most of the dolomite in the Woodford isresedimented but not terrigenous in origin, thenit must be penecontemporaneous. Early forma-tion of dolomite in marine sediment is promotedby hypersaline brine (Zenger, 1972) and by lowconcentrations of dissolved sulfate that developin organic-rich sediments as the result ofmicrobial sulfate reduction (Baker and Kastner,1981). Given the abundance of organic matterand the presence of anhydrite in the Woodford,both are plausible mechanisms for contempo-raneous dolomitization in the Permian Basinduring the Late Devonian.

Areas that have a high ratio of dolomite toquartz (fig. 9f) are found in the central, northern,eastern, and westernmost parts of the basin,suggesting that these were the areas of highestcarbonate production. The highest dolomite/quartz ratio is in the center of the basin (app. C;C2) where very little detrital quartz is found,and the quartz typically is much finer grainedthan dolomite. This observation is furtherevidence that the basin center, which coincideswith the modern-day Central Basin Platform,was bypassed by siliciclastic detritus.

Secondary silica is a common cement inprimary sedimentary structures, such as burrowsand syneresis cracks, where it is associatedlocally with calcite and anhydrite. Burrows andsyneresis cracks are abundant in the northern,central, and eastern basin (fig. 9g) in areas thatwere onlapped by Woodford sediments (forexample, the Northwestern and Eastern Shelves,Central Basin Platform, and western MidlandBasin). They are less abundant or absent in coresfarther east in the Midland Basin. The distribu-tion and association with anhydrite suggest thatthese structures are shallow-water indicatorsformed under dysaerobic conditions above theanoxic zone.

Benthic components are scarce and includetrilobite fragments, brachiopods, and biogenicpellets. Some of the pellets in siltstone and othersassociated with scattered trilobite fragments inshale may be fecal material from a sparsebenthos. However, many are found in blackshale that has parallel laminae and has noburrows or benthic fossils, suggesting that theyoriginated in the upper water column amid athriving, normal marine biota. Most benthicfossils are found in the shelf regions, but biogenicpellets are also common in rocks from theCentral Basin Platform (app. C; C2, C12).

29

Page 34: Stratigraphic Analysis of the - University of Texas at Austin

Depositional SettingPaleogeography

Late Devonian paleogeography of the studyarea (fig. 10) was inferred from the patternsof onlap (pls. 3 through 7) and lithology (fig. 9)described earlier. The widespread, blanketlikedistribution and nearly uniform lithology of theWoodford indicate that the entire region wasone of low relief during the Late Devonian.Major topographic features in the model include(1) the land in the north and northwest repre-senting the Pedernal Massif and Concho Arch,(2) the ancestral Delaware and Val Verde Basins,(3) the shallow Midland Basin, (4) an intrabasinalarchipelago representing the ancestral CentralBasin Platform and Pecos Arch, (5) shallow shelfregions to the north and east representing theancestral Northwestern and Eastern Shelves,(6) a western shelf that had irregular channelsand shoals representing parts of the North-western Shelf and Diablo Platform, and (7) a

land mass to the southwest representing thesouthern part of the Diablo Platform (fig. 10).

The Pedernal Massif and northern ConchoArch represent the southern end of the Trans-continental Arch, which was the dominanttopographic high in the western North Americancraton during the Late Devonian (Poole andothers, 1967; Poole, 1974; Heckel and Witzke,1979). Grain size and composition of Woodfordsiltstones indicate that this arch supplied mostof the terrigenous sediment to the basin andconsequently must have had the highest ele-vations in the study area. The absence of deltasand coarse clastic wedges, however, indicatesthat elevations were not high enough to createan orographic barrier to winds or to introducemajor rainfall, runoff, and clastic influx into thebasin.

The Northwestern and Eastern Shelves andthe Diablo Platform are onlapped by Woodford

Page 35: Stratigraphic Analysis of the - University of Texas at Austin

sediment (pls. 3 through 7), indicating that theywere low-relief expanses of intermediate ele-vation, and that during the Late Devoniantransgression they became shallow-water shelfenvironments that had local channels, scatteredislands, and protected shoals. The western-most outcrop belt is characterized by complexfacies changes (Stevenson, 1945; Laudon andBowsher, 1949; Rosado, 1970), indicating that itcomprised an extensive, low-relief cratonic shelfthat had prominent shoals and channels (Rosado,1970). The southern Diablo Platform may haveremained emergent, but it was not a majorsource of terrigenous sediment (Wright, 1979).

The deepest parts of the Late Devonian epeiricsea coincided with the deepest parts of thepresent-day Delaware and Val Verde Basins(fig. 10), where the thickest and most completeWoodford sections are found (pls. 3 through 7).Gradual changes in well log signatures atformation boundaries in these depocenterssuggest that the Woodford may be conformablewith the bounding formations (pl. 5, C–C´, wells4, 6; pl. 6, D–D ,́ wells 10, 11, 13; pl. 7, E–E ,́ wells16, 17). That the Midland Basin was a topo-graphic depression (fig. 10) is supported by thefollowing evidence: (1) the Woodford thickensand contains all three units toward the basinaxis and (2) the Woodford generally has nobottom features (such as anhydrite-bearingburrows and syneresis cracks), that wouldindicate elevations above the anoxic and sulfate-reducing zones.

The ancestral Central Basin Platform andPecos Arch are shown as a continuous intra-basinal archipelago (fig. 10). Whether the twoactually connected is unknown, but the onlap ofboth structures by the Woodford indicates thatboth were topographically high during the LateDevonian. Lithologic patterns (fig. 9) indicatethat the Central Basin Platform was bypassedby terrigenous sediment, and stratigraphic onlapindicates that bypassing occurred because theplatform was elevated above the surroundingprovinces. The abundance of dysaerobic primarysedimentary structures on the platform (fig. 9g)suggests a shallow-water setting and supportsthis conclusion. Folk (1959) inferred the presenceof an island chain along the platform during

the Early Ordovician on the basis of the abun-dance of feldspar in the Ellenburger Formation.Similarly, the presence of recycled vitrinite inthe black shale lithofacies (fig. 9c) indicates thateroding bedrock existed nearby and that scat-tered islands lay along the platform during LateDevonian eustatic highstand.

PaleotectonicsEllison (1950) recognized anomalously thin

but complete Woodford intervals on structuralhighs along the Central Basin Platform andinterpreted them as evidence of contempo-raneous uplift during Woodford deposition. Pre-Mississippian truncation of the Woodford alongthe ancestral Central Basin Platform (forexample, pl. 4, well 9) and on the NorthwesternShelf (app. B; C5, C9), where the lower unit iswell developed and the upper unit is absent, isfurther evidence of contemporaneous uplift inthese areas. Vertical tectonic adjustments in theLate Devonian most likely reflect reactivation ofbasement structures because truncated sectionsare found along zones of weakness in thebasement and near the major Paleozoic faultsystems (pls. 1, 2) that formed along reactivatedbasement faults (Walper, 1977; Muehlberger,1980; Hills, 1984). In figure 10, contemporaneousvertical movements are illustrated by theschematic representations of normal faults in theDelaware and Val Verde Basins. These faultsrepresent the dominant Paleozoic faults shownin plates 1 and 2.

Epeirogeny in the southern Midcontinentprobably was linked to renewed tectonism alongthe continental margins. The Acadian orogenyproduced highlands (fig. 11) that shed coarseterrigenous elastics toward the craton to formthe Catskill delta (Ettensohn and Barron, 1981;Faill, 1985; Ettensohn, 1987). The Antler orogenyalso produced a rising highland (fig. 11) thatshed coarse elastics into a subsiding forelandbasin (Poole and others, 1967; Poole, 1974).Forces transmitted from the Antler erogenicbelt have been correlated with minor faulting,uplift, and subsidence in New Mexico (Pooleand others, 1967) and are inferred to accountfor Late Devonian epeirogeny in the studyarea.

31

Page 36: Stratigraphic Analysis of the - University of Texas at Austin

PaleoclimateThe paucity of terrestrial organic matter in

the Woodford Formation, including the siltstonelithofacies, suggests that land in the region wasmostly barren, and the absence of coarse-grainedsediments and thick deltaic or fan depositsindicates that the land was low lying and notdrained by large rivers. Furthermore, the pres-ence of anhydrite in primary sedimentarystructures documents hypersalinity within thebasin. Together these observations indicate thatthe Permian Basin was arid during the LateDevonian. An arid paleoclimate and hyper-salinity suggest that some of the dolomite in theWoodford formed in shallow-water evaporiticsettings. Episodic resedimentation by bottomflows would account for the hybrid mixture of

dolomite and quartz grains composing gradedlayers and Bouma divisions.

Arid-climate indicators support a Paleogeo-graphic reconstruction in which the study arealies along the western margin of North Americaat approximately 15 degrees south latitude inthe warm, arid southern trade-wind belt betweenthe wet equatorial doldrums and the wet south-ern temperate zone (Heckel and Witzke, 1979;fig. 11). In this reconstruction the Late Devonianpaleoequator lies along the Antler orogenic beltand the Canadian Rockies from California toAlberta. Other plate tectonic reconstructions ofthe Late Devonian also place the study area atlow southern latitudes in the warm tropics oron the paleoequator (Lowe, 1975; Ettensohn andBarron, 1981; Parrish, 1982).

32

Page 37: Stratigraphic Analysis of the - University of Texas at Austin

PaleoceanographyFeatures characteristic of black shale in the

Woodford, including high organic content,abundant pyrite, and parallel laminae, indicatethat bottom waters were stagnant and anoxicduring deposition. The abundance of pelagicmarine fossils and marine types of organicmatter indicates that surface waters supporteda luxuriant, normal marine biota. Coexistenceof a putrid bottom and fertile surface watersrequires a strongly stratified water column andimplies the presence of a pycnocline (Byers, 1977;Arthur and Natland, 1979; Demaison and Moore,1980; Ettensohn and Barron, 1981; Stanley, 1983;Ettensohn and Elam, 1985; Stein, 1986). The aridclimate and hypersaline indicators imply thata pycnocline formed as a result of the strongdensity contrast between warm, normal-salinitysurface water and cold, somewhat hypersalinebottom water. Anaerobic conditions developedbelow the pycnocline because no vertical mixingwas occurring and because oxygen had beendepleted owing to the high demand created bydecay of the large volume of organic matter.

The abundance of marine organic matter andpelagic fossils indicates that efficient circulationof surface water and continuous resupply ofnutrients characterized the upper part of thewater column. Upwelling off the west andsouthwest coasts of North America during theLate Devonian (Lowe, 1975; Heckel and Witzke,1979; Parrish, 1982) was the most likely sourceof the nutrients. No record exists of large riversdischarging into the basin (that is, deltas or fans)that would indicate a significant, continuousterrestrial source. Published circulation modelssuggest that oceanic surface currents flowingalong the continental margin were divertednorthward and northeastward, carrying up-welled water onto the North American craton(Lowe, 1975; Heckel and Witzke, 1979; Ettensohnand Barron, 1981). The model shown in figure 12suggests that upwelled water moved eastwardinto the basin primarily as counter currents. Inthe southeast trade-wind belt, net flow of surfacewater would have been directed westward outof the basin by the Coriolis force and the Ekmanspiral. The arid climate that produced hyper-

salinity caused net evaporation of surface water,particularly over shallow-water shelves, plat-forms, and shoals. The loss of surface water viawind-driven currents and evaporation wouldhave amplified the negative water balancerequired by eustatic rise, causing inflowingcounter currents to be stronger than outflowingsurface currents.

The model in figure 12 differs from otherpublished models (Lowe, 1975; Heckel, 1977;Demaison and Moore, 1980; Witzke, 1987) in thatthe floor of the basin in this model remainedstagnant and anoxic, receiving sulfide-rich mudthat had parallel laminae, even though netevaporation, local brine production, and nega-tive water balance was occurring. This happenedbecause the circulation pattern developed dur-ing a major marine transgression; therefore,much of the increased volume of water flowingonto the craton can be accounted for by theaddition of hypersaline brine to stagnant bottomwaters. Consequently, dense water graduallyfilled depressions in the epeiric sea without deepcirculation being necessary to maintain waterbalance.

The existence of only dysaerobic (siltstone)and anaerobic (black shale) biofacies in theWoodford Formation indicates that bottomwater became depleted in oxygen soon after theLate Devonian transgression began. Earlyoxygen depletion most likely was related to theearly development of hypersalinity and strongdensity stratification. Dense water accumulatedat the bottom of the water column in topographiclows and probably caused many local pycno-clines to develop during the initial stages oftransgression. Later, at transgressive highstand,a single pycnocline (fig. 12) apparently devel-oped, allowing anaerobic mud, represented bythe black shales of the middle Woodford unit,to accumulate uniformly across the entire region.Dysaerobic bottom indicators found locally inthe black shale on topographic highs (burrows,syneresis cracks, and anhydrite) may recordsome of the small-scale eustatic regressionsdocumented by Johnson and others (1985) andreflect short-term fall of the pycnocline causedby falling sea level.

33

Page 38: Stratigraphic Analysis of the - University of Texas at Austin

Depositional MechanismsBecause the study area was once located in

the tropics (fig. 11), and particularly because theLate Devonian was an epoch of worldwidetransgression and global warming (Johnson andothers, 1985), storms were most likely frequentand geologically significant events (Marsagliaand Klein, 1983; Morton, 1988; Barron, 1989).Frequent storms are therefore the most plausiblemechanism for explaining the generation ofbottom flows. Triggering mechanisms for bot-tom flows include (1) turbid, dense dischargefrom deltas, submarine fans, and rivers in flood,(2) spontaneous slumping of rapidly deposited,unconsolidated sediment, (3) slope failure re-sulting from earthquakes, and (4) sedimentliquefaction and autosuspension during storms(Walker, 1984).

The absence of deltas and submarine fans inthe Woodford precludes the first two mech-anisms. What little turbid flood dischargeentered the basin would not have been denseenough to sink beneath marine or hypersalinebasin water (Drake, 1976; Pierce, 1976). Mostlikely, flood discharge was hypopycnal, or itproduced detached turbidity layers by processessimilar to those that had occurred in modernsubmarine canyons off southern California

(Pierce, 1976) and in the Nile cone and Hellenictrench regions of the Mediterranean Sea(Maldonado and Stanley, 1978; Stanley andMaldonado, 1981). Deposition from turbid,muddy plumes would not produce gradedlayers or Bouma sequences but could yield thevarvelike laminae (Pierce, 1976; Stanley, 1983)characteristic of the black shale lithofacies inthe Woodford.

Earthquakes associated with epeirogenicmovements probably triggered some bottomflows, but the subtlety of structural displacementduring the Late Devonian indicates that thesemovements probably were weak and infrequent.Furthermore, bottom flows starting in shallowwater would be diverted along the pycnoclinein strongly stratified seas (Pierce, 1976; Stanley,1983), unless they entrained brine from restrictedhypersaline basins, shelves, or shoals (Arthurand Natland, 1979).

Storms, rather than earthquakes, probablywere the most frequent and powerful agents ofsediment transport in the warm Late Devoniantropics. They can account for both the indis-criminate mixing of siliciclastic and dolomitegrains and the generation of bottom flows thatpersisted into basinal depocenters. In modernseas, storms can disrupt density stratification

34

Page 39: Stratigraphic Analysis of the - University of Texas at Austin

(Mooers, 1976a, b), a condition that could mini-mize flow detachment and promote sustainedbottom flows. It is probable that such a processhappened in Late Devonian times as well. Stormwinds and surge would flush shallow-water,hypersaline environments and give rise to verydense bottom flows consisting of sediment-ladenbrine. Briny bottom flows would maintain theirintegrity below the pycnocline even in stronglystratified basins.

Evidence indicates that bottom flows peri-odically disturbed anoxia that existed beneaththe pycnocline. In black shales, burrows arecommonly confined to graded layers and Boumadivisions, indicating that the bottom was brieflyinhabited by organisms after sediment depo-sition. Bottom flows originating in shallow,aerobic or dysaerobic environments apparentlyentrained enough oxygen to sustain a temporarybenthic population. However, oxygen wasquickly depleted by the meager fauna, decay oforganic matter, and absence of oxygen resupply.And because bottom oxygenation was short-lived, anaerobic conditions quickly returned,killing the few allochthonous organisms. Bur-rowed layers in the Cretaceous Mowry Shale(Davis and others, 1989) and the DevonianChattanooga Shale (Potter and others, 1982) havebeen similarly explained, and entrainment ofoxygen and benthic organisms in turbidity cur-rents apparently occurred in modern sedimentsin the Santa Barbara Basin (Sholkovitz andSoutar, 1975).

Basal siltstones in proximal shelf and basinenvironments (app. B; C5, C9, C11) consist ofvertically stacked siltstone-shale couplets, docu-menting episodic deposition from bottom flows.The greater numbers and thicknesses of silt-stones in the deepest parts of the Delaware,Midland, and Val Verde Basins indicate thatthese depocenters were locations where bottomflows, initiated in various parts of the basin,finally converged. The high frequency of bottomflows in basinal depocenters implies that basinaxes were dysaerobic more often than were distalshelves, slopes, and platforms. Thus, the lowerconcentrations of organic matter in the basins(fig. 9e) can be attributed to the combined effectsof dilution by clastic sediment and destruction

by oxidation, aerobic microbes, and the tem-porary benthos.

Synopsis of Depositional HistoryWoodford deposition began when the sea

drowned marine embayments in what are nowthe deepest parts of the Delaware and Val VerdeBasins and advanced over a subaerially erodedand dissected terrane composed mostly of car-bonate rocks of Ordovician to Middle Devonianage. A broad epeiric sea formed that had irreg-ular bottom topography and scattered, low-reliefland masses. The basin lay in the arid midtropicssurrounded by lands that supported little vege-tation and few rivers. Oceanic water from anarea of coastal upwelling flowed into the ex-panding epeiric sea and maintained a thriving,normal marine biota in the upper levels of thewater column. Net evaporation locally producedhypersaline brines, and strong density stratifi-cation developed that restricted vertical circu-lation. The basin quickly became dysaerobic andthen anaerobic as sea level continued to rise.Once oxygen was eliminated from the bottom,sulfide-rich mud began to accumulate. Risingsea level and persistent oceanographic andclimatic patterns allowed anaerobic mud depo-sition to continue slowly during the rest of theLate Devonian Epoch. Frequent storms andoccasional earthquakes triggered bottom flowsthat supplied silty mud to proximal shelves anddeep basin troughs and caused much resedi-mentation throughout the basin. Tectonic stressarising from the Antler orogeny initiatedepeirogenic movements throughout the regionand caused contemporaneous movements alongreactivated basement faults.

Woodford deposition probably ended becausesea level stabilized or dropped and ocean-ographic patterns changed, thus halting thestrong net flow of ocean water onto the cratonand forcing deep circulation through most ofthe basin. Glauconite and calcified benthic epi-fauna accumulated on the floor of the epeiricsea, marking a change in bottom conditions fromanaerobic to dysaerobic and locally aerobic andrecording the improved vertical circulationthrough most of the basin.

35

Page 40: Stratigraphic Analysis of the - University of Texas at Austin

Petroleum PotentialThe Woodford Formation is currently gen-

erating oil in the Midland Basin, Central BasinPlatform, and Eastern and Northwestern Shelvesand is currently generating gas in the Delawareand Val Verde Basins. Thermal maturity of theWoodford Formation was deduced from thedepth and Ro data in appendix D and the depthversus Ro log-normal relationship derived forthe Woodford in the Anadarko Basin (Cardott,1989). Oil generation in the Woodford occursbetween R0 values of 0.5 and 1.3 percent(Cardott, 1989) at depths between 6,000 and13,000 ft in the Permian Basin. These depthscorrespond to depths below sea level of approx-imately 4,000 to 10,000 ft in the region east ofthe Central Basin Platform and 2,000 to 9,000 ftin the Delaware Basin and regions to the west(fig. la; pl. 1). Condensate and wet-gas genera-tion occurs between R0 values of 1.3 and2.0 percent (Cardott, 1989) at depths between13,000 and 18,000 ft common only in theDelaware and Val Verde Basins. These depthscorrespond to depths below sea level of approx-imately 9,000 to 14,000 ft in the region west ofthe Central Basin Platform and south of thePecos Arch (fig. la; pl. 1). Dry gas is generatedbetween R0 values of 2.0 and 5.0 percent atdepths between 18,000 and 26,000 ft (Cardott,1989), or at depths below sea level of 14,000 to22,000 ft in the Delaware and Val Verde Basins(fig. 1a; pl. 1).

Commercial production of hydrocarbonsfrom the Woodford is possible in areas wherethe formation is highly fractured. The fracturedUpper Devonian shales (Ohio, Chattanooga,Antrim, Bakken, and Woodford) that producegas in the Appalachian and Michigan Basins andoil in the Williston and Ardmore Basins illustratethe commercial potential and provide appro-priate geological models for exploration in thePermian Basin. In West Texas and southeasternNew Mexico, optimum drilling targets are thesiltstones and radiolarian cherts because theyare competent lithologies that are the most likelyto maintain open fracture systems. Areas thathave the greatest density of major faults are themost prospective: these include the Central BasinPlatform, southernmost Midland Basin, andparts of the Northwestern Shelf (fig. la; pls. 1,2). Production may be possible from the well-developed basal siltstone in the northern partof the Midland Basin and adjacent NorthwesternShelf (for example, app. B; C5, C11 in Cochranand Gaines Counties, Texas). Although faultsare uncommon there, commercial productioncould be established in zones where porosityhas been enhanced or permeability can beartificially stimulated. Gas undoubtedly is pres-ent in siltstones and fractured shales in theDelaware and Val Verde Basins; however, drill-ing depths would make costs prohibitive inmost places.

SummaryThe Woodford Formation is an organic-rich

petroleum source rock that has long beenrecognized as an important marker unit becauseof its black shales, anomalously high radio-activity, widespread distribution, and strati-graphic position between carbonates. TheWoodford is mostly Late Devonian in age andis stratigraphically equivalent to the Devo-nian black shales (for example, Chattanooga,Ohio, Antrim, New Albany, Bakken, Exshaw,and Percha) that are present in many NorthAmerican basins. At most localities, the

Woodford overlies a major regional uncon-formity and is diachronous.

In the Permian Basin, the Woodford is thickest(661 ft) in the Delaware Basin depocenter andlocally is absent from structural highs on theCentral Basin Platform and Pecos Arch. Struc-tural relief in the subsurface is 20,000 ft; itdeveloped primarily during the late Paleozoicas a response to erogenic activity in the OuachitaFold Belt.

Two lithofacies, black shale and siltstone,compose the Woodford. The black shale exhibits

36

Page 41: Stratigraphic Analysis of the - University of Texas at Austin

varvelike parallel laminae, abundant pyrite,very high radioactivity, and high concentrationsof marine organic matter (mean = 4.5 ± 2.6 wt %TOC). It is the most widely distributed anddistinctive rock type in the formation. Siltstoneis a hybrid of silt-sized quartz and dolomitegrains and exhibits discontinuous or disruptedstratification, graded layers, fine-grained Boumasequences, and moderately high radioactivity.It is restricted to deep basin and proximal shelfsettings and is commonly the basal unit. Onthe basis of lithology and stratigraphic posi-tion, basal siltstone is correlated with the Onateand Canutillo Formations in New Mexico andWest Texas, the Misener and Sylamore Sand-stones in Oklahoma and Arkansas, and the IvesBreccia Member of the Houy Formation inCentral Texas. The black shale lithofacies iscorrelated with the Sly Gap and Percha For-mations in the west and the Doublehorn Shaleand phosphatic members of the Houy Formationin Central Texas. Black shale is mostly pelagicand represents an anaerobic biofacies, whereassiltstone was deposited by bottom flows andcomprises a dysaerobic biofacies. Upwardtransition from basal siltstone to black shalelocally records the worldwide marine trans-gression that occurred during the Late Devonian.

The Woodford onlaps Paleozoic structuresflanking the Midland, Delaware, and Val VerdeBasins, indicating that all of the major structuralprovinces in the modern-day Permian Basin hadtopographic expression in the Late Devonian.The blanketlike geometry and nearly uniformlithology, however, indicate that the regionwas one of low relief. The increased size andabundance of siliciclastic grains (quartz, mus-covite, feldspar) and wood fragments in thenorthern part of the basin show that the PedernalMassif and northern Concho Arch were theprincipal source areas of terrigenous sediment.In contrast, most dolomite formed contempo-raneously on distal platforms and shelves inhighly reduced, low-sulfate mud or restrictedmarine environments. Resedimentation of dolo-mite grains and mixing with siliciclastics wereaccomplished by bottom flows.

Woodford black shale records widespreadbottom stagnation and anoxia during depositionand a strongly density-stratified water column.

High concentrations of marine organic matterand siliceous pelagic micro-organisms in theshale indicate high biological productivity insurface waters supported mainly by dynamicupwelling. Episodes of hypersalinity, docu-mented by the presence of anhydrite in bur-rows and syneresis cracks, suggest an aridpaleoclimate and indicate that density stratifi-cation was caused, at least partly, by accu-mulation of hypersaline bottom water.

The plate tectonic reconstruction most con-sistent with an arid paleoclimate and dynamicupwelling places the study area on the westernmargin of North America in the dry tropics near15 degrees south latitude. In this setting,southeasterly trade winds and the Ekman spiralwould push surface waters westward towardthe open ocean and upwelled oceanic watereastward onto the craton as counter currents.The negative water balance required for marinetransgression would be amplified by flow intothe basin replacing water lost by evaporation.

This circulation model accounts for the largesupply of nutrients needed to support highbiological productivity in the upper part of thewater column of the epeiric sea. Furthermore,the low-latitude paleogeography and LateDevonian global warming imply frequent trop-ical storms and suggest that the bottom flowsthat caused the deposition of hybrid quartz/dolomite siltstones were storm generated.

The end of Woodford deposition coincidedwith the end of the Late Devonian eustatic rise.Bottom oxygenation, recorded as accumulationsof glauconite and calcitic benthic fossils, indicatesthat new oceanographic conditions includeddeep circulation in most of the basin. Thestabilization or fall of sea level would haveended the strong net flow of ocean water con-taining upwelled nutrients onto the craton andforced deep circulation to maintain waterbalance.

The Woodford Formation is now in the oilwindow in the Midland Basin, Central BasinPlatform, and Eastern and NorthwesternShelves, and it is in the gas window in theDelaware and Val Verde Basins. Commercialproduction of hydrocarbons is possible fromintervals that are highly fractured, but optimumdrilling targets are siltstone and radiolarian chert

37

Page 42: Stratigraphic Analysis of the - University of Texas at Austin

beds in densely faulted regions, such as theCentral Basin Platform, southernmost MidlandBasin, and parts of the Northwestern Shelf.

Development of reserves in unusual geologicalsettings such as the Woodford Formation in thePermian Basin undoubtedly will be required tomeet future demands for petroleum. These

reserves can be discovered through compre-hensive studies, similar to the present report,that integrate stratigraphic, petrologic, andgeochemical data. Such studies can help predictthe location and lithology of unconventional oiland gas reservoirs that are inherently difficultto find.

AcknowledgmentsThis study, funded by the Bureau of Economic

Geology, at The University of Texas at Austin,was begun while the author was a ResearchFellow at the Bureau on sabbatical from TulsaUniversity (TU). Peter K. Krynine, projectResearch Assistant, helped compile the welllog data, describe core, and conduct the fieldwork; his efforts are gratefully acknowledged.David V. LeMone (The University of Texas atEl Paso [UTEP]) shared his thoughts on thestratigraphy and origin of the Percha, providedrelevant UTEP theses, suggested appropriateoutcrops for study, and accompanied the authorto Bishop Cap, New Mexico. Paul C. Franks (TU)read parts of the manuscript and offered helpfulsuggestions. The manuscript benefited greatlyfrom critical reviews by S. C. Ruppel, W. B.Ayers, Jr., E. H. Guevara, W. A. Ambrose, andA. R. Scott, all at the Bureau.

Cores were supplied by Shell Oil Company,Mobil Oil Corporation, and the Exxon Company.

Vitrinite reflectance and visual kerogen analyseswere performed by Robert Littlejohn, KerogenPrep, Inc., Tulsa, Oklahoma, and TOC wasdetermined by Geochem Laboratories, Inc.,Houston, Texas.

Maps and cross sections (pls. 1 through 7)were drafted at the Bureau of Economic Geologyby Margaret D. Koenig and Kerza Prewitt, underthe supervision of Richard L. Dillon. The textfigures were drafted in Tulsa: the fence diagramby Sharon S. Scott and the lithology logs andPaleogeographic block diagram by Herbert S.Scott. Final modifications were made at theIndiana Geological Survey (IGS) and the Bureau.Drafts of the manuscript were typed by MargaretAndrews (TU) and Marilyn DeWees (IGS). Wordprocessing and typesetting were done at theBureau by Susan Lloyd under the supervisionof Susann Doenges. Technical editing was byTucker F. Hentz. Margaret L. Evans designedthe report. Lana Dieterich was the editor.

38

Page 43: Stratigraphic Analysis of the - University of Texas at Austin

ReferencesAigner, Thomas, 1982, Calcareous tempestites: storm-dominated

stratification in Upper Muschelkalk limestones (Middle Trias,SW-Germany), in Einsele, Gerhard, and Seilacher, Adolf, eds.,Cyclic and event stratification: Berlin, Springer-Verlag,p. 180–198.

——————— 1984, Storm depositional systems, in Friedman,G. M., Neugebauer, H. J., and Seilacher, Adolf, eds., Lecturenotes in earth sciences, v. 3: New York, Springer-Verlag,174 p.

Amsden, T. W., 1975, Hunton Group (Late Ordovician, Silurian,and Early Devonian) in the Anadarko Basin of Oklahoma:Oklahoma Geological Survey Bulletin 121, 214 p.

——————— 1980, Hunton Group (Late Ordovician, Silurian,and Early Devonian) in the Arkoma Basin of Oklahoma:Oklahoma Geological Survey Bulletin 129, 136 p.

Amsden, T. W., Caplan, W. M., Hilpman, P. L., McGlasson,E. H., Rowland, T. L., and Wise, O. A., 1967, Devonian of thesouthern Midcontinent area, United States, in Oswald, D. H.,ed., International symposium on the Devonian System:Calgary, Canada, Alberta Society of Petroleum Geologists,v.l, p. 913–932.

Amsden, T. W., and Klapper, Gilbert, 1972, Misener sandstone(Middle-Upper Devonian), north-central Oklahoma:American Association of Petroleum Geologists Bulletin,v. 56, no. 12, p. 2323–2334.

Arthur, M. A., and Natland, J. H., 1979, Carbonaceous sedimentsin the North and South Atlantic: the role of salinity in stablestratification of Early Cretaceous basins, in Talwani, Manik,Hay, William, and Ryan, W. B. P., eds., Deep drilling resultsin the Atlantic Ocean: continental margins and paleo-environment: American Geophysical Union, Maurice EwingSeries 3, p. 375–401.

Baker, P. A., and Kastner, Miriam, 1981, Constraints on theformation of sedimentary dolomite: Science, v. 213, no. 4504,p. 214–216.

Barron, E. J., 1989, Severe storms during Earth history: GeologicalSociety of America Bulletin, v. 101, no. 5, p. 601–612.

Bolton, Keith, Lane, H. R., and LeMone, D. V., 1982, Symposiumon the paleoenvironmental setting and distribution of theWaulsortian facies: El Paso Geological Society and TheUniversity of Texas at El Paso, 202 p.

Bouma, A. H., 1962, Sedimentology of some flysch deposits:Amsterdam, Elsevier, 168 p.

Broadhead, R. R, Kepferle, R. C, and Potter, P. E., 1982,Stratigraphic and sedimentologic controls of gas in shale—example from Upper Devonian of northern Ohio: AmericanAssociation of Petroleum Geologists Bulletin, v. 66, no. 1,p. 10–27.

Burrowes, O. G., and Krause, F. F., 1987, Overview of theDevonian System: subsurface of Western Canada Basin, inKrause, F. F., and Burrowes, O. G., eds., Devonian lithofaciesand reservoir styles in Alberta, 13th CSPG Core Conferenceand Display and Second International Symposium on theDevonian System: Calgary, Canada, Canadian Society ofPetroleum Geologists, p. 1–20.

Byers, C. W., 1977, Biofacies patterns in euxinic basins: a generalmodel, in Cook, H. E., and Enos, Paul, eds., Deep-water

39

carbonate environments: Society of Economic Paleontologistsand Mineralogists Special Publication No. 25, p. 5–17.

Cardott, B. J., 1989, Thermal maturation of the Woodford Shalein the Anadarko Basin, in Johnson, K. S., ed., Anadarko BasinSymposium, 1988: Oklahoma Geological Survey Circular 90,p. 32–46.

Cloud, P. E., Barnes, V. E., and Hass, W. H., 1957, Devonian-Mississippian transition in central Texas: Geological Societyof America Bulletin, v. 68, no. 7, p. 807–816.

Cluff, R. M., 1980, Paleoenvironment of the New Albany ShaleGroup (Devonian-Mississippian) of Illinois: Journal ofSedimentary Petrology, v. 50, no. 3, p. 767–780.

Cluff, R. M., Reinbold, M. L., and Lineback, J. A., 1981, TheNew Albany Shale Group of Illinois: Illinois State GeologicalSurvey Circular 518, 83 p.

Comer, J. B., and Hinch, H. H., 1987, Recognizing andquantifying expulsion of oil from the Woodford Formationand age-equivalent rocks in Oklahoma and Arkansas:American Association of Petroleum Geologists Bulletin,v. 71, no. 7, p. 844–858.

Davis, H. R., Byers, C. W., and Pratt, L. M., 1989, Depositionalmechanisms and organic matter in Mowry Shale(Cretaceous), Wyoming: American Association of PetroleumGeologists Bulletin, v. 73, no. 9, p. 1103–1116.

Demaison, G. J., and Moore, G. T., 1980, Anoxic environmentsand oil source bed genesis: American Association ofPetroleum Geologists Bulletin, v. 64, no. 8, p. 1179–1209.

Dow, W. G., 1977, Kerogen studies and geological interpretations:Journal of Geochemical Exploration, v. 7, no. 2, p. 77–79.

Drake, D. E., 1976, Suspended sediment transport and muddeposition on continental shelves, in Stanley, D. J., and Swift,D. J. P., eds., Marine sediment transport and environmentalmanagement: New York, John Wiley, American GeologicalInstitute, p. 127–158.

Duncan, D. C, and Swanson, V. E., 1965, Organic-rich shale ofthe United States and world land areas: U.S. GeologicalSurvey Circular 523, 30 p.

Dutton, S. P., Goldstein, A. G., and Ruppel, S. C, 1982, Petroleumpotential of the Palo Duro Basin, Texas Panhandle: TheUniversity of Texas at Austin, Bureau of Economic GeologyReport of Investigations No. 123, 87 p.

Ellison, S. P., Jr., 1950, Subsurface Woodford black shale, WestTexas and southeast New Mexico: University of Texas,Austin, Bureau of Economic Geology Report of InvestigationsNo. 7, 20 p.

Ettensohn, F. R., 1987, Rates of relative plate motion during theAcadian orogeny based on the spatial distribution of blackshales: Journal of Geology, v. 95, no. 4, p. 572–582.

Ettensohn, F. R., and Barron, L. S., 1981, Depositional model forthe Devonian-Mississippian black shales of North America:a paleoclimatic-paleogeographic approach, in Roberts, T. G.,ed., Geological Society of America Cincinnati ’81 Field TripGuidebooks, v. II: Economic Geology, Structure, p. 344–361.

Ettensohn, F. R., and Ham, T. D., 1985, Defining the nature andlocation of a Late Devonian-Early Mississippian pycnoclinein eastern Kentucky: Geological Society of America Bulletin,v. 96, no. 10, p. 1313–1321.

Page 44: Stratigraphic Analysis of the - University of Texas at Austin

Ettensohn, F. R., Fulton, L. P., and Kepferle, R. C, 1979, Useof scintillometer and gamma-ray logs for correlationand stratigraphy in homogeneous black shales: GeologicalSociety of America Bulletin, v. 90, part I, p. 421–423; part n,p. 828–849.

Ewing, T. E., 1991, Tectonic map of Texas: The University ofTexas at Austin, Bureau of Economic Geology, scale 1:750,000.

Faill, R. T., 1985, The Acadian orogeny and the Catskill delta, inWoodrow, D. L, and Sevon, W. D., eds., The Catskill delta:Geological Society of America Special Paper 201, p. 15–37.

Folk, R. L., 1959, Thin-section examination of pre-SimpsonPaleozoic rocks, in Barnes, V. E., ed., Stratigraphy of the pre-Simpson Paleozoic subsurface rocks of Texas and southeastNew Mexico, v. 1: Austin, University of Texas PublicationNo. 5924, p. 95–130.

Folk, R. L., and Pittman, J. S., 1971, Length-slow chalcedony: anew testament for vanished evaporites: Journal ofSedimentary Petrology, v. 41, no. 4, p. 1045–1058.

Francis, B. M., 1988, Petrology and sedimentology of theDevonian Misener Formation, north-central Oklahoma:University of Tulsa, Master’s thesis, 176 p.

Freeman, Tom, and Schumacher, Dietmar, 1969, Qualitativepre-Sylamore (Devonian-Mississippian) physiography delin-eated by onlapping conodont zones, northern Arkansas:Geological Society of America Bulletin, v. 80, no. 11,p.2327–2334.

Frezon, S. E., and Jordan, Louise, 1979, Oklahoma, in Craig, L. G,and Connor, C. W., coordinators, Paleotectonic investigationsof the Mississippian System in the United States, part I,Introduction and regional analyses of the Mississippiansystem: U.S. Geological Survey Professional Paper 1010-1,p. 146–159.

Galley, J. E., 1958, Oil and geology in the Permian Basin ofTexas and New Mexico, in Weeks, L. G., ed., Habitat of oil:Tulsa, Oklahoma, American Association of PetroleumGeologists, p. 395–446.

Click, E. E., 1979, Arkansas, in Craig, L. C., and Connor, C. W.,coordinators, Paleotectonic investigations of the MississippianSystem in the United States, part I, Introduction and regionalanalyses of the Mississippian System: U.S. Geological SurveyProfessional Paper 1010-H,p. 125–145.

Graves, R. W., 1952, Devonian conodonts from the CaballosNovaculite: Journal of Paleontology, v. 26, no. 4, p. 610–612.

Ham, W. E., 1969, Regional geology of the Arbuckle Mountains,Oklahoma, in Ham, W. E., ed., Geology of the ArbuckleMountains: Oklahoma Geological Survey Guidebook 17,p. 5–21.

Ham, W. E., and Wilson, J. L., 1967, Paleozoic epeirogeny andorogeny in the central United States: American Journal ofScience, v. 265, p. 332–407.

Harlton, B. H., 1956, The Harrisburg Trough, Stephens and CarterCounties, Oklahoma, in Hicks, I. C., Westheimer, J. M.,Thomlinson, C. W., Putman, D. M., and Selk, E. L., eds.,Petroleum geology of southern Oklahoma, a symposium,v. 1; sponsored by the Ardmore Geological Society: Tulsa,American Association of Petroleum Geologists, p. 135–143.

Hass, W. H., 1951, Age of Arkansas Novaculite: AmericanAssociation of Petroleum Geologists Bulletin, v. 35, no. 12,p. 2526–2541.

Hass, W. H., and Huddle, J. W., 1965, Late Devonian and EarlyMississippian age of the Woodford Shale in Oklahoma, as

determined from conodonts: U.S. Geological SurveyProfessional Paper 525-D, p. D125–D132.

Hayes, M. O., 1967, Hurricanes as geological agents, south Texascoast: American Association of Petroleum Geologists Bulletin,v. 51, no. 6, p. 937–942.

Heckel, P. H., 1977, Origin of phosphatic black shale facies inPennsylvanian cyclothems of Mid-Continent North America:American Association of Petroleum Geologists Bulletin,v. 61, no. 7, p. 1045–1068.

Heckel, P. H., and Witzke, B. J., 1979, Devonian worldpalaeogeography determined from distribution of carbonatesand related lithic palaeoclimatic indicators, in House, M. R.,Scrutton, C. T., and Bassett, M. G., eds., The Devonian System:Special Papers in Palaeontology No. 23, p. 99–123.

Hein, J. R., and Parrish, J. T., 1987, Distribution of siliceousdeposits in space and time, in Hein, J. R., ed., Siliceoussedimentary rock-hosted ores and petroleum: New York,Van Nostrand Reinhold, p. 10–57.

Hills, J. M., 1984, Sedimentation, tectonism, and hydrocarbongeneration in Delaware Basin, West Texas and southeasternNew Mexico: American Association of Petroleum GeologistsBulletin, v. 68, no. 3, p. 250–267.

Hoenig, M. A., 1976, Stratigraphy of the “Upper Silurian” and“Lower Devonian,” Permian Basin, West Texas: TheUniversity of Texas at El Paso, Master’s thesis, 81 p.

Horak, R. L, 1985, Tectonic and hydrocarbon maturation historyin the Permian Basin: Oil and Gas Journal, v. 83, no. 21,p. 124–129.

Huffman, G. G., 1958, Geology of the flanks of the Ozark uplift:Oklahoma Geological Survey Bulletin No. 77,281 p.

Huffman, G. G., and Starke, J. M., 1960, Noel shale in-northeastern Oklahoma: Oklahoma Geology Notes, v. 20,p. 159–163.

Hunt, J. M., 1979, Petroleum geochemistry and geology: SanFrancisco, Freeman, 617 p.

Johnson, J. G., Klapper, Gilbert, and Sandberg, C. A., 1985,Devonian eustatic fluctuations in Euramerica: GeologicalSociety of America Bulletin, v. 96, no. 5, p. 567–587.

Jones, T. S., and Smith, H. M., 1965, Relationships of oilcomposition and stratigraphy in the Permian Basin of WestTexas and New Mexico, in Young, Addison, and Galley,J. E., eds., Fluids in subsurface environments: Tulsa, AmericanAssociation of Petroleum Geologists, p. 101–224.

Keroher, G. C., and others, 1966, Lexicon of geologic names ofthe United States for 1936–1960: U.S. Geological SurveyBulletin 1200, part 3, P-Z, p. 4292–4293.

King, P. B., King, R. E., and Knight, J. B., 1945, Geology of theHueco Mountains, El Paso and Hudspeth Counties, Texas:U.S. Geological Survey, Oil and Gas InvestigationsPreliminary Map 36.

Kottlowski, F. E., 1963, Paleozoic and Mesozoic strata ofsouthwestern and south-central New Mexico: New MexicoBureau of Mines and Mineral Resources Bulletin 79,p. 24–32.

Landis, E. R., 1962, Uranium and other trace elements inDevonian and Mississippian black shales in the centralMidcontinent area: U.S. Geological Survey Bulletin 1107-E,p. 289–336.

Laudon, L. R., and Bowsher, A. L., 1949, Mississippian formationsof southwestern New Mexico: Geological Society of AmericaBulletin, v. 60, no. 1, p. 1–88.

40

Page 45: Stratigraphic Analysis of the - University of Texas at Austin

Leggett, J. K., 1980, British lower Paleozoic black shales andtheir paleo-oceanographic significance: Geological Society ofLondon Journal, v. 137, pt. 2, p. 139–156.

LeMone, D. V., 1971, General stratigraphy of the FranklinMountains, in Robledo Mountains, New Mexico, FranklinMountains, Texas, 1971 Field Conference: Society of EconomicPaleontologists and Mineralogists Permian Basin SectionPublication 71-13, p. ix.

Leventhal, J. S., 1981, Pyrolysis gas chromatography-massspectrometry to characterize organic matter and itsrelationship to uranium content of Appalachian Devonianblack shales: Geochimica et Cosmochimica Acta, v. 45, no. 6,p. 883–889.

Lindberg, F. A., ed., 1983, Correlation of Stratigraphic Units ofNorth America (COSUNA) Project, Southwest/SouthwestMid-Continent Region: Tulsa, American Association ofPetroleum Geologists.

Lloyd, E. R., 1949, Pre-San Andres stratigraphy and oil-producingzones in southeastern New Mexico: New Mexico Bureau ofMines and Mineral Resources Bulletin 29, 87 p.

Lowe, D. R., 1975, Regional controls on silica sedimentation inthe Ouachita system: Geological Society of America Bulletin,v. 86, no. 8, p. 1123–1127.

Lucia, F. J., 1971, Lower Paleozoic history of the western DiabloPlatform, West Texas and south central New Mexico, inRobledo Mountains, New Mexico, Franklin Mountains,Texas, 1971 Field Conference: Society of EconomicPaleontologists and Mineralogists Permian Basin SectionGuidebook Publication 71-13, p. 174–214.

Maldonado, Andres, and Stanley, D. J., 1978, Nile conedepositional processes and patterns in the Late Quaternary,in Stanley, D. J., and Kelling, Gilbert, eds., Sedimentation insubmarine canyons, fans, and trenches: Stroudsburg,Pennsylvania, Dowden, Hutchinson and Ross, p. 239–257.

Marsaglia, K. M., and Klein, G. deV., 1983, The paleogeographyof Paleozoic and Mesozoic storm depositional systems:Journal of Geology, v. 91, no. 4, p. 117–142.

Mapel, W. J., Johnson, R. B., Bachman, G. O., and Varnes,K. L., 1979, Southern Midcontinent and southern RockyMountain region, in Paleotectonic investigations of theMississippian System in the United States, part I, Introduc-tion and regional analyses of the Mississippian System:U.S. Geological Survey Professional Paper 1010-J, p. 160–187.

McGlasson, E. H., 1967, The Siluro-Devonian of West Texasand southeast New Mexico, in Oswald, D. H., ed., Inter-national symposium on the Devonian System: Calgary,Canada, Alberta Society of Petroleum Geologists, v. 2,p. 935–948.

Meissner, F. F., 1978, Petroleum geology of the BakkenFormation, Williston Basin, North Dakota and Montana:Montana Geological Society 24th Annual Conference, 1978Williston Basin Symposium, p. 207–227.

Mooers, C. N. K., 1976a, Introduction to the physicaloceanography and fluid dynamics of continental margins,in Stanley, D. J., and Swift, D. J. P., eds., Marine sedimenttransport and environmental management: New York,JohnWiley,p.7–21.

——————— 1976b, Wind-driven currents on the continentalmargin, in Stanley, D. J., and Swift, D. J. P., eds., Marinesediment transport and environmental management: NewYork, John Wiley, p. 29–52.

Morton, J. P., 1985, Rb-Sr dating of diagenesis and source age ofclays in Upper Devonian black shales of Texas: GeologicalSociety of America Bulletin, v. 96, no. 8, p. 1043–1049.

Morton, R. A., 1981, Formation of storm deposits by wind-forced currents in the Gulf of Mexico and the North Sea, inNio, S. D., Shattenhelm, R. T. E., and Van Weering, T. C. E.,eds., Holocene marine sedimentation in the North Sea Basin:International Association of Sedimentologists SpecialPublication 5, p. 385–396.

———————1988, Nearshore responses to great storms, inClifton, H. E., ed., Sedimentologic consequences of convulsivegeologic events: Geological Society of America Special Paper229, p. 7–22.

Muehlberger, W. R., 1980, Texas lineament revisited: NewMexico Geological Society Guidebook, 31st Field Confer-ence, Trans-Pecos Region, p. 113–121.

Munn, J. K, 1971, Breedlove field, Martin County, Texas:American Association of Petroleum Geologists Bulletin,v. 55, no. 3, p. 403–411.

Nelson, C. H., 1982, Modern shallow-water graded sand layersfrom storm surges, Bering Shelf: a mimic of Bouma sequencesand turbidite systems: Journal of Sedimentary Petrology,v. 52, no. 2, p. 537–545.

Odin, G. S., and Letolle, Rene, 1980, Glauconitization andphosphatization environments: a tentative comparison, inBentor, Y. K., ed., Marine phosphorites—geochemistry,occurrence, genesis: Society of Economic Paleontologistsand Mineralogists Special Publication No. 29, p. 227–237.

Park, D. E., and Croneis, Carey, 1969, Origin of Caballos andArkansas Novaculite formations: American Association ofPetroleum Geologists Bulletin, v. 53, no. 1, p. 94–111.

Parrish, J. T., 1982, Upwelling and petroleum source beds, withreference to Paleozoic: American Association of PetroleumGeologists Bulletin, v. 66, no. 6, p. 750–774.

Parrish, J. T., and Barron, E. J., 1986, Paleoclimates and economicgeology: Society of Economic Paleontologists and Miner-alogists Short Course Notes No. 18, 162 p.

Pashin, J. C., and Ettensohn, F. R., 1987, An epeiric shelf-to-basin transition: Bedford-Berea sequence, northeasternKentucky and south-central Ohio: American Journal ofScience, v. 287, p. 893–926.

Pedersen, T. F., and Calvert, S. E., 1990, Anoxia vs. productivity:what controls the formation of organic-carbon-rich sedimentsand sedimentary rocks: American Association of PetroleumGeologists Bulletin, v. 74, no. 4, p. 454–466.

Peirce, F. L., 1962, Devonian cavity fillings in subsurface Siluriancarbonates of West Texas: American Association of PetroleumGeologists Bulletin, v. 46, no. 1, p. 122–124.

Pierce, J. W., 1976, Suspended sediment transport at the shelfbreak and over the outer margin, in Stanley, D. J., and Swift,D. J. P., eds., Marine sediment transport and environmentalmanagement: New York, John Wiley, p. 437–458.

Piper, D. J. W., 1978, Turbidite muds and silts on deep sea fansand abyssal plains, in Stanley, D. J., and Kelling, Gilbert, eds.,Sedimentation in submarine canyons, fans, and trenches:Stroudsburg, Pennsylvania, Dowden, Hutchinson and Ross,p. 163–176.

Pittenger, H. A., 1981, Provenance, depositional environmentand diagenesis of the Sylamore sandstone in northeasternOklahoma and northern Arkansas: University of Tulsa,Master’s thesis, 109 p.

41

Page 46: Stratigraphic Analysis of the - University of Texas at Austin

Poole, F. G., 1974, Flysch deposits of Antler foreland basin,western United States, in Dickinson, W. R., ed., Tectonics andsedimentation: Society of Economic Paleontologists andMineralogists Special Publication No. 22, p. 58–82.

Poole, R G., Baars, D. L, Drewes, Harald, Hayes, P. T., Ketner,K. B., McKee, E. D., Teichert, Curt, and Williams, J. S., 1967,Devonian of the southwestern United States, in Oswald,D. H., ed., International symposium on the DevonianSystem, v. 1: Calgary, Canada, Alberta Society of Petro-leum Geologists, p. 879–912.

Potter, P. E., Maynard, J. B., and Pryor, W. A., 1982, Appalachiangas bearing Devonian shales: statements and discussions:Oil and Gas Journal, v. 80, no. 4, p. 290–318.

Pratt, L. M., 1984, Influence of paleoenvironmental factors onpreservation of organic matter in middle CretaceousGreenhorn Formation, Pueblo, Colorado: American Associ-ation of Petroleum Geologists Bulletin, v. 68, no. 9,p. 1146–1159.

Pray, L. C., 1961, Geology of the Sacramento Mountainsescarpment, Otero County, New Mexico: New MexicoBureau of Mines and Mineral Resources Bulletin 35, 144 p.

Rhoads, D. C., and Morse, J. W., 1971, Evolutionary and ecologicsignificance of oxygen-deficient marine basins: Lethaia, v. 4,no. 4, p. 413–428.

Roen, J. B., 1984, Geology of the Devonian black shales of theAppalachian basin: Organic Geochemistry, v. 5, no. 4,p. 241–254.

Rosado, R. V., 1970, Devonian stratigraphy of south-centralNew Mexico and far West Texas: The University of Texas atEl Paso, Master’s thesis, 108 p.

Ruppel, S. C., 1985, Stratigraphy and petroleum potential ofpre-Pennsylvanian rocks, Palo Duro Basin, Texas Panhandle:The University of Texas at Austin, Bureau of EconomicGeology Report of Investigations No. 147, 81 p.

Schieber, Jiirgen, 1987, Storm-dominated epicontinental clasticsedimentation in the mid-Proterozoic Newland Formation,Montana, U.S.A.: Neues Jahrbuch fur Geologic undPalaontologie Monatshefte, v. 7, p. 417–439.

Sholkovitz, Edward, and Soutar, Andrew, 1975, Changes in thecomposition of the bottom water of the Santa Barbara Basin:effect of turbidity currents: Deep-Sea Research, v. 22, no. 1,p. 13–21.

Schopf, T. J. M., 1983, Paleozoic black shales in relation tocontinental margin upwelling, in Thiede, John, and Suess,Erwin, eds., Sedimentary records of ancient coastal upwelling:coastal upwelling, its sediment record, part B: New York,Plenum Press, p. 579–596.

Stanley, D. J., 1983, Parallel laminated deep-sea muds andcoupled gravity flow-hemipelagic settling in the Mediter-ranean: Smithsonian Contributions to the Marine SciencesNo. 19, 19 p.

Stanley, D. J., and Maldonado, Andres, 1981, Depositionalmodels for fine-grained sediment in the western Hellenictrench, eastern Mediterranean: Sedimentology, v. 28, no. 2,p. 273–290.

42

Stein, Ruediger, 1986, Organic carbon and sedimentation rate—further evidence for anoxic deep-water conditions in theCenomanian/Turonian Atlantic Ocean: Marine Geology,v. 72, no. 3/4, p. 199–209.

Stevenson, F. V., 1945, Devonian of New Mexico: Journal ofGeology, v. 53, p. 217–245.

Stow, D. A. V., and Piper, D. J. W., 1984, Deep-water fine-grained sediments: facies models, in Stow, D. A. V., andPiper, D. J. W., eds., Fine-grained sediments: deep-waterprocesses and facies: Geological Society of London,Special Publication No. 15, p. 611–646.

Sutherland, P. K., and Manger, W. L., 1979, Comparison ofOzark shelf and Ouachita basin facies for UpperMississippian and Lower Pennsylvanian series in easternOklahoma and western Arkansas, in Sutherland, P. K., andManger, W. L., eds., Ozark and Ouachita shelf-to-basintransition, Oklahoma-Arkansas: Norman, Oklahoma,Oklahoma Geological Survey Guidebook 19, p. 1–13.

Swanson, V. E., 1960, Oil yield and uranium content of blackshales: U.S. Geological Survey Professional Paper 356-A,44 p.

——————— 1961, Geology and geochemistry of uranium inmarine black shales—a review: U.S. Geological SurveyProfessional Paper 356-C, 112 p.

Swift, D. J. P., Figueiredo, A. G., Freeland, G. L., and Oertel,G. P., 1983, Hummocky cross-stratification and megaripples:a geological double standard: Journal of SedimentaryPetrology, v. 53, no. 4, p. 1295–1317.

Taff, J. A., 1902, Geological atlas of the United States: U.S.Geological Survey Atoka Folio No. 79.

Tissot, B. P., Demaison, G. J., Masson, Peter, Delteil, J. R., andCombaz, Andre, 1980, Paleoenvironment and petroleumpotential of Middle Cretaceous black shale in Atlantic basins:American Association of Petroleum Geologists Bulletin, v. 64,no. 12, p. 2051–2163.

Tissot, B. P., and Welte, D. H., 1984, Petroleum formation andoccurrence, 2d ed.: New York, Springer-Verlag, 699 p.

Walker, R. G., 1984, Turbidites and associated coarse clasticdeposits, in Walker, R. G., ed., Facies models, 2d ed.:Geoscience Canada, Reprint Series 1, p. 171–188.

_______ 1985, Geological evidence for storm transportationand deposition on ancient shelves, in Tillman, R. W., Swift,D. J. P., and Walker, R. G., eds., Shelf sands and sandstonereservoirs: Society of Economic Paleontologists and Mineral-ogists Short Course Notes 13, p. 243–302.

Walper, J. L., 1977, Paleozoic tectonics of the southern marginof North America: Gulf Coast Association of GeologicalSocieties Transactions, v. 27, p. 230–241.

Witzke, B. J., 1987, Models for circulation patterns inepicontinental seas applied to Paleozoic facies of NorthAmerica craton: Paleoceanography, v. 2, no. 2, p. 229–248.

Wright, W. F., 1979, Petroleum geology of the Permian Basin:Midland, Texas, West Texas Geological Society, 98 p.

Zenger, D. H., 1972, Dolomitization and uniformitarianism:Journal of Geological Education, v. 20, no. 3, p. 107–124.

Page 47: Stratigraphic Analysis of the - University of Texas at Austin

APPENDICES

Page 48: Stratigraphic Analysis of the - University of Texas at Austin
Page 49: Stratigraphic Analysis of the - University of Texas at Austin
Page 50: Stratigraphic Analysis of the - University of Texas at Austin
Page 51: Stratigraphic Analysis of the - University of Texas at Austin
Page 52: Stratigraphic Analysis of the - University of Texas at Austin
Page 53: Stratigraphic Analysis of the - University of Texas at Austin
Page 54: Stratigraphic Analysis of the - University of Texas at Austin
Page 55: Stratigraphic Analysis of the - University of Texas at Austin
Page 56: Stratigraphic Analysis of the - University of Texas at Austin
Page 57: Stratigraphic Analysis of the - University of Texas at Austin
Page 58: Stratigraphic Analysis of the - University of Texas at Austin
Page 59: Stratigraphic Analysis of the - University of Texas at Austin
Page 60: Stratigraphic Analysis of the - University of Texas at Austin
Page 61: Stratigraphic Analysis of the - University of Texas at Austin
Page 62: Stratigraphic Analysis of the - University of Texas at Austin
Page 63: Stratigraphic Analysis of the - University of Texas at Austin
Page 64: Stratigraphic Analysis of the - University of Texas at Austin
Page 65: Stratigraphic Analysis of the - University of Texas at Austin
Page 66: Stratigraphic Analysis of the - University of Texas at Austin