Anda di halaman 1dari 25

High-resolution stratigraphy AUTHORS

and facies architecture Michael D. Fairbanks ~ Apache


Corporation, 2000 Post Oak Boulevard, Suite
100, Houston, Texas 77056-4400; michael.
of the Upper Cretaceous fairbanks@apachecorp.com

(Cenomanian–Turonian) Eagle Michael D. Fairbanks is a petroleum geologist


for Apache Corporation. He graduated from
The University of Texas at Austin in 2012 with an
Ford Group, Central Texas M.S. in geology, and he received his B.S. in
geology from Brigham Young University, Utah,
Michael D. Fairbanks, Stephen C. Ruppel, and in 2010. His interests include sedimentological
Harry Rowe and depositional controls in unconventional
reservoir systems and chemostratigraphy of
source rock intervals.

Stephen C. Ruppel ~ Bureau of Economic


ABSTRACT Geology, The University of Texas at Austin,
Rock-based studies of the Eagle Ford Group of Central Texas University Station, Box X, Austin, Texas
demonstrate that mudrock deposition is more complicated than 78713-8924; stephen.ruppel@beg.utexas.edu
previously supposed. X-ray diffraction, x-ray fluorescence, total Stephen C. Ruppel has over 30 years of
organic carbon (TOC), and log data collected from eight cores and experience in sedimentological and
two outcrops demonstrate that bottom-current reworking and geochemical characterization of carbonate and
planktonic productivity are primary depositional controls, acting mudrock reservoir systems. He holds a Ph.D.
from the University of Tennessee as well as B.S.
independently from eustatic forcing. Central Texas Eagle Ford
and M.S. degrees from the University of Illinois
facies include (1) massive argillaceous mudrock, (2) massive and the University of Florida, respectively. He
foraminiferal calcareous mudrock, (3) laminated calcareous for- has published more than 200 scientific papers
aminiferal lime mudstone, (4) laminated foraminiferal wacke- and reports and is director of the Bureau’s
stone, (5) cross-laminated foraminiferal packstone–grainstone, Mudrocks Systems Research Laboratory.
(6) massive bentonitic claystone, and (7) nodular foraminiferal
Harry Rowe ~ Bureau of Economic
packstone–grainstone. High degrees of lateral facies variability, Geology, The University of Texas at Austin,
characterized by pinching and swelling of units, lateral facies University Station, Box X, Austin, Texas
changes, truncations, and locally restricted units, are observed 78713-8924; harry.rowe@beg.utexas.edu
even at small lateral scales (50 ft [15 m]). At 10 mi (16 km) and Harry Rowe is a geologist at the Bureau of
greater lateral spacings, core and geochemical data signifi- Economic Geology, Austin, Texas. His main
cantly underestimate intraformational facies variability. Ap- interest is in developing chemostratigraphic
proximately 73% of units can be successfully correlated across a records of paleoceanographic change.
distance of 500 ft (152 m), 35% are traceable across 1 mi
(1.6 km), and only 16% of beds are correlative across 10 mi ACKNOWLEDGMENTS
(16 km). Geochemical proxies (enrichment in molybdenum
and other trace elements) indicate that maximum anoxia oc- This paper comprises parts of the thesis of
Michael D. Fairbanks at The University of
curred within the Bouldin Member despite being composed of
Texas at Austin, under the direction of
the most calcareous and high-energy facies. Comparison of Stephen C. Ruppel and William L. Fisher.
total gamma ray (GR) logs to computed GR logs is requisite, Funding for this project came from the
because GR alone may provide misleading determination of industry members of the Mudrock Systems
Research Laboratory Consortium at the
Bureau of Economic Geology. Industry
Copyright ©2016. The American Association of Petroleum Geologists. All rights reserved. members include Anadarko, Apache, BP,
Manuscript received October 8, 2014; provisional acceptance January 7, 2015; revised manuscript received Chesapeake, Chevron, Cima, Cimarex,
August 10, 2015; final acceptance December 7, 2015. ConocoPhilips, Cypress, Devon, Encana, EOG,
DOI:10.1306/12071514187

AAPG Bulletin, v. 100, no. 3 (March 2016), pp. 379–403 379


EXCO, Husky, Marathon, Pangaea, Penn facies, TOC content, depositional environment, and sequence
Virginia, Penn West, Pioneer, Shell, StatOil, stratigraphic implications.
Texas American Resources, The
Unconventionals, US EnerCorp, Valence, and
YPF. The Jackson School of Geosciences and
the Bureau of Economic Geology provided INTRODUCTION
the opportunity, resources, and facilities for
the study. Thanks go to Timothy Kearns, Brett The Upper Cretaceous Eagle Ford Group has long been recog-
Huffman, Robert Nikirk, and Jessica Buckles nized as an important source rock for productive reservoirs
for their assistance in acquisition and analysis throughout Texas. Heightened industry focus on the Eagle Ford is
of x-ray fluorescence data while at The a result of recent discoveries of producible unconventional pe-
University of Texas at Arlington, to Necip troleum resources in this emerging play. However, little has been
Guven of The University of Texas at San
published on the facies and facies heterogeneities within this
Antonio, and to Henry Francis of the
Kentucky Geological Survey, Lexington, for mixed-carbonate–clastic mudrock system. A rock-based study is
integrated analysis of x-ray diffraction data. fundamental to understanding the controls, types, and scales of
Thanks also go to Ryan Harbor, who laid the inherent heterogeneities, which have implications for enhanced
foundation, and Greg Frébourg and Bob comprehension of the Eagle Ford Group and other mixed-
Loucks, whose continual insights contributed carbonate–clastic mudrock depositional systems worldwide.
to the progression of this work. This study uses a unique data set of densely distributed cores
to constrain the controls on inherent facies heterogeneities and
variabilities, which helps to answer the following questions
regarding mudrock systems: What is the continuity of units
expressed by wireline log response? What causes variable
production response, and how can production be optimized
across large play areas? What controls the distribution of res-
ervoir properties?
The primary objectives of this study are to (1) define Eagle
Ford lithofacies present in the Austin, Texas, area; (2) determine
lithofacies continuity on various scales, along with the processes
that control intraformational heterogeneities; (3) explore the ef-
fectiveness of elemental abundance data in identifying lithological
variations within the Eagle Ford system; (4) use geochemical data
to refine lithofacies interpretations; and (5) evaluate the applic-
ability of total gamma ray (GR) logs in defining facies and rock
properties.
The eight cores and two outcrops of this rare data set span a
nearly 11 mi (18 km) transect, with seven of the cores located
within a 2 mi (3 km) area, providing a uniquely high-resolution
perspective for mudrock stratigraphy and facies continuity
(Figure 1). Energy-dispersive x-ray fluorescence (XRF), x-ray
diffraction (XRD), Rock-Eval total organic carbon (TOC), and
thin-section synthesis further enhance this study.

REGIONAL GEOLOGY

The southwesternmost expression of the Grenville orogeny is the


1.15-1 Ga Llano uplift in Central Texas, resulting in vast ex-
posures of deformed crystalline basement, consisting of gneisses,
schists, and metavolcanic rocks (Culotta et al., 1992). An extension

380 Eagle Ford Stratigraphy and Facies Architecture


Figure 1. Location map of the study area showing cores and outcrops in the Austin, Texas, area and cores used for regional correlation.
The Eagle Ford outcrop belt spans the Texas–Mexico border, passes through the study area, and continues across the Texas–Oklahoma
border and into Arkansas. Eagle Ford outcrop belt adapted from Barnes (1992). 204 = BT-204; 221 = BT-221 core; 222 = BT-222-PTPZ core;
301 = BO-301-PTPZ core; 302 = BO-302-PT core; 500 = BI-500-PT core; 514 = BI-514-PTPZ core; ACC = ACC 1 core; Blumberg = J. W.
Blumberg 1-B core; Brechtel = W. Brechtel 1 core; Hendershot = C. J. Hendershot 1 core; Orts = H. P. Orts 2 core; Schauer = F. T.
Schauer et al. 1 core; SH = state highway; St = street.

Fairbanks et al. 381


Figure 2. Paleogeographic map of Texas at the Cenomanian–Turonian time boundary. The study area lies on the northeastern flank of
the San Marcos arch, which separates the Maverick Basin from the East Texas Basin. The Sligo and Stuart City shelf margins form a
physiographic break toward the downdip extent of the Eagle Ford. Modified from Ruppel et al. (2012).

of the uplift, the San Marcos arch, trends southeast– was a low-lying subaerial terrain, receiving and sup-
northwest (Dravis, 1980; Young, 1986) and, in the plying little sediment, similar to modern-day Florida
Early Cretaceous, was a minor topographic high (Tyler (Young, 1986). The study location in Austin lies on the
and Ambrose, 1986; Figure 2) where a large carbonate northeastern flank of the San Marcos arch (Figure 2).
platform developed (Donovan and Staerker, 2010). Thick progradational carbonate packages com-
During most of the Cretaceous, the San Marcos arch prising the Comanche shelf developed during the

Figure 3. Generalized cross


section through the Cretaceous
Gulf of Mexico Basin. The Eagle
Ford overlies thick progradational
packages of the Lower Cretaceous
Comanche shelf. Modified from
Ruppel et al. (2012).

382 Eagle Ford Stratigraphy and Facies Architecture


Table 1. Description of Data Set, Identifying Cores, and Outcrop Locations

Austin Area Cores*


Core Number Name in Study Eagle Ford Depths (ft [m]) County

BT-222-PTPZ 222 51–88 (16–27) Travis


BI-500-PT 500 55–94 (17–29) Travis
BI-514-PTPZ 514 52–89 (16–27) Travis
BT-204 204 80–116 (24–35) Travis
BO-302-PT 302 52–90 (16–27) Travis
BO-301-PTPZ 301 51–88 (16–27) Travis
BT-221 221 52–88 (16–27) Travis
ACC 1 ACC 80–125 (24–38) Travis

Regional Cored Wells

Wells Name in Study Eagle Ford Depths (ft [m]) County API Operator
C. J. Hendershot 1 Hendershot 4734–4774 (1443–1455) Caldwell 4217730218 Tesoro Petroleum
W. Brechtel 1 Brechtel 3280–3315 (1000–1010) Wilson 4249330208 Prairie Producing Co.
H. P. Orts 2 Orts 7684–7757 (2342–2364) Gonzales 4217730203 Transocean Oil, Inc.
F. T. Schauer et al. 1 Schauer 8093–8159 (2467–2487) Gonzales 4217730394 Geological Res Corp.
J. W. Blumberg 1-B Blumberg 4175–4225 (1273–1288) Wilson 4218730532 Prairie Producing Co.
Burkland 1 Burkland 935–977 (285–298) Caldwell 4205534144 Vista Energy Corp.

*Outcrop: West Bouldin Creek, approximate coordinates: N30°1599.88860 and W97°45941.31660; Walnut Creek, approximate coordinates: N30°24928.25940 and
W97°42931.79520.

Early Cretaceous (Figure 3). During the Hauterivian which contributed a substantial clay mineral com-
Stage of the Early Cretaceous, the Sligo shelf margin ponent within the Eagle Ford facies. Figure 2
developed (Salvador and Muñeton, 1989), forming summarizes the paleogeographic setting near the
a raised rim shelf margin profile (Galloway, 2008; Cenomanian–Turonian boundary. Eagle Ford dep-
Figure 3). Flooding during the Aptian Stage initiated osition was immediately followed by Coniacian–
a landward shift in reef development, forming the Santonian Austin Chalk deposition and subsequently
Stuart City shelf margin (Salvador and Muñeton, the rest of the Upper Cretaceous Series strata (Hentz
1989; Phelps, 2011). Until the Cenomanian, rudist–coral and Ruppel, 2010; Hentz et al., 2014; Figure 3).
communities acted as constructers and formed baffles
to normal wave action, which formed the nearly con-
tinuous carbonate reef of the Stuart City rimming the STUDY AREA
Gulf of Mexico (Sohl et al., 1991; Scott, 2010; Figure 2).
A major transgression of the Comanchean In Central Texas, the Eagle Ford outcrop belt passes
shelf in the middle-to-late Cenomanian initiated the through Travis County and directly through the city
deposition of the Eagle Ford Group (Sohl et al., 1991; of Austin (Figure 1). An outcrop located along West
Phelps, 2011; Figure 3). To the northeast, sedimentation Bouldin Creek forms the type locality for the
in the Woodbine delta system resulted in deposition of Bouldin Member (Adkins and Lozo, 1951; Jiang,
the coeval Tuscaloosa and Woodbine Formations in 1989) and has been described in detail by previous
a fluvial and deltaic setting (Sohl et al., 1991; Dubiel authors (Feray and Young, 1949; Adkins and Lozo,
et al., 2010), as well as incised-valley-fill, nearshore 1951; Pessagno, 1969; Young, 1977; Liro et al., 1994;
marine, and wave-dominated-delta deposits (Hentz Lundquist, 2000). Another outcrop in north Austin is
et al., 2014). During the Cretaceous, much of the encountered along a cutbank of Walnut Creek, south
riverborne detritus included illite (Pratt, 1984), of Park Bend Road (Figure 1).

Fairbanks et al. 383


Located between the West Bouldin Creek and as proxies for calcite, clay minerals, and quartz, re-
Walnut Creek outcrops, 11 cores comprising the entire spectively) to aid in facies definition. Trace elements
Eagle Ford interval were also used in this study (e.g., molybdenum (Mo), uranium (U), vanadium
(Figure 1). One core, ACC 1, was recovered during (V)) help to evaluate levels of oxygenation during
the development of an Edwards aquifer monitoring deposition (Tribovillard et al., 2006). Elemental
well in north Austin. Also, 10 additional geotechnical abundances of U, thorium (Th), and potassium (K)
cores form a tightly spaced, 1-mi (1.6-km)-long can be used to create the equivalent of a total GR
transect along Waller Creek in Austin (Figure 1).
for purposes of comparing core data to borehole
To obtain a regional perspective and evaluate the
log data through the formula
regional correlatability of the Eagle Ford in the Austin
study, this work includes cored subsurface wells, ðU8Þ + ðTh4Þ + ðK16Þ = GR
which were described by Harbor (2011; Figures 1, 2).
where U is the uranium content (ppm), Th is the
Two cross-section lines span the San Marcos arch, one
thorium content (ppm), and K is the weight percent
extending to the southwest from Austin across the
arch, and the other extending to the south along the of potassium (Doveton and Merriam, 2004). An
dip of the arch (Figure 1). The subsurface Eagle Ford equivalent computed GR (CGR) log can be calcu-
extends in a dip section from the shallow surface and lated by omitting the U, which ensures that the signal
outcrop belt to the Late Cretaceous shelf margin, with is representative of the clay mineral content and re-
depths reaching 14,000 ft (4267 m) (Harbor, 2011). flects lithological changes. To replicate the appear-
Table 1 summarizes the data set used in the current ance of downhole logs, discrete XRF data points are
study and the cores studied by Harbor (2011). smoothed and plotted.
Samples for thin-section, organic geochemistry,
and XRD analysis were removed from the intact core
METHODS AND DATA and prepared and processed by National Petrographic
Services, GeoMark Research Ltd., and Necip Guven,
Handheld energy-dispersive XRF is quickly be- respectively. Sample locations were selected based on
coming an industry standard for acquiring inorganic representative facies identification and distributed
geochemical data from cores and cuttings of mudrock throughout the cores. A thorough discussion of the
intervals. Elemental calibrations involved reference methods involved in acquiring these data are provided
materials from five internationally accepted standards, by Harbor (2011) and Fairbanks (2012).
SDO-1, SGR-10, SCo-1, GBW-0717, and SARM-41,
along with 86 reference materials from diverse mu-
drock systems, including the Woodford, Smithwick, RESULTS
Barnett, Eagle Ford, and Ohio Shale (Kearns, 2011;
Rowe et al., 2012). Facies
All analyses were undertaken using a Bruker AXS
Tracer III-V XRF unit, equipped with an Rh x-ray Although deceptively homogenous at the core-face
tube. Major element analyses were conducted at 15 scale because of their grain size and dark color, mu-
kV, with the beam and detector under vacuum. Trace drocks are extremely heterogeneous; this has important
element analyses were conducted at 40 kV, with an implications for depositional environment and sedi-
Al–Ti–Cu beam filter. Core samples for XRF analysis ment delivery processes. The following seven distinctive
were taken at a 1 ft (31 cm) interval and were ana- facies are identified within the Eagle Ford group: (1)
lyzed by placing the flat slab side down on the massive argillaceous mudrock, (2) massive fora-
nose of the instrument. Raw sample XRF spectra miniferal calcareous mudrock, (3) laminated cal-
were calibrated using proprietary Bruker software careous foraminiferal lime mudstone, (4) laminated
(Rowe et al., 2012). foraminiferal wackestone, (5) cross-laminated for-
Dominant mineral abundances are quickly de- aminiferal packstone–grainstone, (6) massive
termined from major element composition (e.g., bentonitic claystone, and (7) nodular foramini-
calcium (Ca), aluminum (Al), and silicon (Si) serve feral packstone–grainstone. A summary of Eagle

384 Eagle Ford Stratigraphy and Facies Architecture


Table 2. Summary Table of Central Texas Eagle Ford Facies and Associated Attributes

Facies Name Sedimentary Structures Mineralogy Dominant Fauna Color Occurrence (Formation) Thickness
Massive argillaceous Massive 53% clay minerals, Rare globigerinid Dark gray to Pepper Shale 4–6 ft (1.2–1.8 m)
mudrock 32% calcite, foraminifera medium-dark gray
17% quartz

Massive foraminiferal Massive 50% calcite, 38% Globigerinid foraminifera, Medium-dark gray Unnamed unit, South 1–5 ft (0.3–1.5 m)
calcareous mudrock clay minerals, inoceramid, and Bosque Formation
14% quartz pelecypod fragments

Laminated calcareous Very thin planar 62% calcite, 24% Globigerinid foraminifera, Medium gray Unnamed unit, South 1–5 ft (0.3–1.5 m)
foraminiferal lime laminations, scours clay minerals, inoceramid, and Bosque Formation
mudstone 8% quartz pelecypod fragments

Laminated foraminiferal Very thin planar–subplanar 76% calcite, 10% Globigerinid foraminifera, Medium gray with Bouldin Member 0.5–1 ft (15–30 cm)
wackestone laminations, slump folds, clay minerals, inoceramid, and white-to–light-gray
fine ripple laminations, 7% quartz pelecypod fragments bands
scours

Cross-laminated Thin subplanar and 86% calcite, 5% Globigerinid foraminifera, Light-to-medium gray Unnamed unit, 0.5–6 in. (1–15 cm)
foraminiferal cross-laminations, scours quartz, 3% clay inoceramid, and Bouldin Member
packstone– grainstone minerals pelecypod fragments

Massive bentonitic Massive 69% smectite, Rare globigerinid Whitish gray to Primarily Bouldin 0.5–6 in. (1–15 cm)
claystone 12% kaolinite, foraminifera medium rust orange Member
8% illite

Nodular foraminiferal Massive, planar and 84% calcite, Globigerinid foraminifera, Light-to-medium gray Bouldin Member, 1–6 in. (2–15 cm)
packstone– grainstone cross-laminated 13% clay minerals, inoceramid, and unnamed unit
2% quartz pelecypod fragments

Fairbanks et al.
385
Figure 4. Composite strati-
graphic chart reviewing nomen-
clature used in the Central Texas
Eagle Ford intervals, compared
with those used in West Texas
and near Waco, Texas. This study
adopts the names of Pepper
Shale, unnamed unit of the Lake
Waco Formation, Bouldin Mem-
ber of the Lake Waco Formation,
and South Bosque Formation.

Ford facies and their associated attributes is found from the scope of this study, previous studies have
in Table 2. shown that much of the fine silt and clay-sized car-
bonate in Eagle Ford rocks is composed of coccolith
debris. The depositional environment of this facies is
Massive Argillaceous Mudrock Facies
interpreted to have been in anoxic marine conditions
Restricted to the basal Pepper Shale of the Eagle Ford
below storm weather wave base, as suggested by the
Group (Figure 4), this facies is dark-to–medium-dark
paucity of benthonic fauna, bioturbation, and wave-
gray in core (Figure 5), weathers recessively in out-
generated structures.
crop, and displays a distinctively smooth or soapy
texture (Table 2). It splits into thin sheets along
planar as well as irregular partings, with common Massive Foraminiferal Calcareous Mudrock Facies
ammonite impressions in the partings. The mineral Massive foraminiferal calcareous mudrock is the
composition, as revealed by XRD and XRF data, is most common facies in the Eagle Ford succes-
primarily clay minerals (average 53%, range sion in Central Texas, and it is found in the
32%–68%), which in decreasing abundance include unnamed member as well as the South Bosque
illite, kaolinite, illite and smectite mixed, and Formation (Figure 4). Medium-to-dark gray in
smectite. Calcite and dolomite constitute an average core, this facies is dominantly structureless and
of 32% (range 3%–60%). Although not apparent weathers recessively (Figure 5). Compared with

386 Eagle Ford Stratigraphy and Facies Architecture


Figure 5. Core photographs,
(A) (C) thin-section photomicrographs,
and summaries of characteristic
features of massive argillaceous
mudrock facies (A and B) and
massive foraminiferal calcareous
mudrock facies (C and D). (A)
Slab photograph showing the
massive, dark-gray fissile char-
acter. (B) Thin-section photo-
micrograph depicting the
fine-grained nature and rare
globigerinid foraminifera. (C)
Slab photo showing the massive,
medium-dark-gray character.
(D) Thin-section photomicro-
graph showing fine-grained na-
ture and abundant globigerinid
foraminifera, inoceramid, and
bivalve shells.

(B) (D)

the massive argillaceous mudrock facies, this suggest that the depositional environment was
facies is coarser grained and more calcareous, in oxygen-poor marine conditions below storm
and it contains a higher quantity of planktonic weather wave base.
(forams) and benthonic (inoceramid and bivalve
fragments) fauna (Table 2). The XRF and XRD Laminated Calcareous Foraminiferal Lime Mudstone Facies
data indicate that this facies is composed of car- This facies is composed of very thinly planar-to-
bonate (average 50%, range 36%–68%) and clay subplanar laminated medium-dark gray (Figure 6),
minerals (average 39%, range 21%–67%). The calcareous mudrock-to-lime mudstone that is present
dark-gray color, presence of low-oxygen ben- in the unnamed unit and the South Bosque Formation
thonic fauna, and lack of sedimentary structures (Figure 4). Laminations are primarily composed of

Fairbanks et al. 387


grain-sized globigerinid foraminifera, with lesser Laminated Foraminiferal Wackestone Facies
amounts of inoceramid and other bivalve fragments. Restricted to the Bouldin Member (Figure 4), this
Calcite and dolomite content from XRD analysis facies is characterized by higher calcite (average
average 62% (range 57%–73%), whereas clay mineral 76%, range 75%–80%) and abundant globigerinid
content measures an average of 24% (range 18%–31%) foraminifera compared with previously described
(Table 2). These rocks are interpreted to have facies (Figure 6; Table 2). Although generally a
been deposited below storm weather wave base wackestone, these rocks locally grade into foramini-
where suspension settling supplied planktonic com- feral packstones or grainstones. Based on the high
ponents and bottom-current-winnowing–created planar degree of current-induced structures (very thin planar
laminations. and subplanar laminations, fine ripple laminations,

Figure 6. Core photographs,


thin-section photomicrographs,
(A) (D)
and summaries of characteristic
features of laminated calcareous
foraminiferal lime mudstone fa-
cies (A–C) and laminated fora-
miniferal wackestone (D and E).
(A) Slab photograph showing the
laminated, medium-gray charac-
ter. (B) Thin-section photo show-
ing fine lamination and abundant
globigerinid foraminifera. (C)
Thin-section photomicrograph
showing erosional scouring. (D)
Slab photograph showing the fine
ripple laminations and medium-
gray character. (E) Thin-section (B)
photomicrograph showing dis-
turbed bedding, abundant globi-
gerinid foraminifera, and
inoceramid fragments. (E)

(C)

388 Eagle Ford Stratigraphy and Facies Architecture


Figure 7. Core photographs,
thin-section photomicrographs,
(A) (C) and summaries of characteristic
features of cross-laminated fora-
miniferal packstone–grainstone
(A and B) and massive bentonitic
claystone facies (C and D). (A)
Slab photograph showing the
cross-laminated and light-gray
nature. (B) Thin-section photo-
micrograph showing the abun-
dant globigerinid foraminifera,
inoceramid fragments, bioclasts,
pyrite, and grain-rich texture. (C)
Slab photograph showing the
structureless, poorly lithified na-
ture. (D) Thin-section photo-
(D) micrograph showing rare
globigerinid foraminifera and
(B) fine-grained texture.

scours) and slump folds, this facies is interpreted to mostly fine sand-sized globigerinid foraminifera
have been deposited below storm weather wave base tests, as well as highly abraded inoceramid and
on a gradually dipping seafloor with bottom-current other bivalve fragments (Figure 7). Cross-laminated
reworking. sedimentary structures and pervasive scouring sug-
gest that this facies records the highest energy of
Cross-Laminated Foraminiferal Packstone–Grainstone deposition within the study area. Based on sedi-
Facies mentary structures and paleoredox proxies from
Highly calcareous (XRD average 86%, range 82%–97% XRF, the depositional environment of this fa-
calcite) (Table 2), this facies is encountered solely cies is interpreted to have been an oxygen-poor
in the Bouldin Member (Figure 4) and contains sediment–water interface in a marine basin where

Fairbanks et al. 389


bottom-current reworking was the prominent sedi- These rocks are present primarily in the Bouldin
ment dispersal mechanism and sediment supply was Member (Figure 4), with few occurrences in the
accelerated by stimulated organic productivity in the unnamed unit and South Bosque Formation. Several
photic zone. studies have interpreted these rocks as volcanic ash
deposits (Jiang 1989; Liro et al., 1994; Dawson 1997)
Massive Bentonitic Claystone Facies potentially sourced from northern Mexico and the
In cores and outcrop, thin (0.5–6 in. [1–15 cm]) Balcones igneous province (Pierce, 2014).
bentonitic claystone beds display a characteristically
recessive and poorly lithified nature (Figure 7; Table 2),
are structureless, and contain only trace foramin- Nodular Foraminiferal Packstone–Grainstone Facies
ifera. The XRD analysis reveals that clay minerals, Resistant in outcrop, this facies comprises horizontal
dominated by smectite and kaolinite, are their dom- lenses that exhibit oblate ellipsoidal cross sections
inant constituent, averaging 91% (range 88%–93%). and differential compaction in adjacent mudstones

Figure 8. Core photographs,


thin-section photomicrographs, (A) (B)
and summary of characteristic
features of nodular foraminiferal
packstone–grainstone (A–D). (A
and B) Slab photographs de-
picting characteristic expression
in core, where (A) displays in-
clined bedding at the terminus of
a nodule, and (B) displays a
gradational contact. (C) Outcrop
photograph depicting the oblate
ellipsoid geometry (12-in. [31-cm]
hammer for scale). (D) Thin-
section photomicrograph showing
abundant globigerinid foramin-
ifera and cemented nature.
(C)

(D)

390 Eagle Ford Stratigraphy and Facies Architecture


(Figure 8). The primary allochem within this totaling 10 ft (3 m) of thickness in the ACC 1 core
facies is globigerinid foraminifera (Table 2). This (Table 1).
facies differs from the cross-laminated foraminiferal The Bouldin Member of the Lake Waco For-
packstone–grainstone facies largely in that it is laterally mation (Figure 4) averages 10–12 ft (3.5–4.0 m) of
discontinuous in outcrop and even at the core scale. thickness in the study area. The Bouldin is more
In core, it is difficult to differentiate between this calcite rich than other Central Texas Eagle Ford Units
facies and cross-laminated foraminiferal packstone– (Figure 4), the calcite being primarily composed of
grainstone facies. These rocks are most abundant planktonic foraminifera, which is consistent with the
in the Bouldin Member, but they are also present findings of recent investigators (Denne et al., 2014).
in the unnamed unit (Figure 4). Frébourg et al. Dominant facies include cross-laminated foramini-
(2016) proposed that similar nodular foraminiferal feral packstones–grainstones, laminated foraminiferal
packstone–grainstones in West Texas were deposited wackestones, massive bentonitic claystones, and
by bottom currents as sediment waves and dunes nodular foraminiferal packstones–grainstones (Table 2).
and that they represent reworked accumulations of A particularly interesting aspect of the bentonitic
planktonic debris. claystone facies is their strong association with the
cross-laminated foraminiferal packstone–grainstones
Stratigraphy and nodular foraminiferal packstones–grainstones. Ex-
cept for a few isolated beds in the underlying unnamed
In Central Texas, the Eagle Ford Group is typically unit and the overlying South Bosque Formation, ben-
divided into two or three units (Figure 4). A recent tonitic claystones are largely confined to the Bouldin
study (Fairbanks, 2012) has proposed that the section Member. Also confined primarily to the Bouldin
can be subdivided into four units: a basal Pepper Shale, Member are the laminated foraminiferal wackestones
an unnamed unit that has been informally referred to and the cross-laminated foraminiferal packstone–
as the Waller member (Fairbanks, 2012; Denne and grainstone facies (Table 2). The strong association
Breyer, in press), an overlying Bouldin Member, and between abundant planktonic foraminifera and
an uppermost South Bosque Formation (Figure 4). bentonite claystones suggests that these deposits are
The Pepper Shale (Figure 4) is a recessive, dark genetically related. Recent studies have connected
gray, argillaceous claystone, comprised solely of the surface water nutrient enrichment to volcanic erup-
massive argillaceous mudrock facies (Figure 6; tions (Duggen et al., 2007, 2010; Jones and Gislason,
Table 2). This interval is relatively thin; it is 4–6 ft 2008; Langmann et al., 2010). It is proposed that
(0.6–1.0 m) thick in the study area (Figure 4). This nutrients introduced into the ocean water system
highly siliceous and argillaceous unit is interpreted by by volcanic eruptions led to higher productivity
recent authors as distal deltaic and prodelta facies within the oxic zone and stimulated planktonic
(Hentz et al., 2014), suggesting time equiva- blooms that led to higher rates of carbonate
lence to the Woodbine section (Adkins and Lozo, sedimentation.
1951; Childs et al., 1988; Dawson et al., 1993) facies. The contact between the Bouldin Member and
Overlying the Pepper Shale is an unnamed the underlying argillaceous foraminiferal mudrock
member of the Lake Waco Formation (Figure 4). of the unnamed unit is picked as the base of the first
Visually, the basal contact of this unit is gradational bed of cross-laminated foraminiferal packstone–
and subtle. The contact is marked by the transition grainstone facies. The upper contact is defined by
to a gritty calcareous mudrock composed of massive the last occurrence of foraminiferal packstone–
foraminiferal calcareous mudrock and laminated grainstone.
calcareous foraminiferal lime mudstone (Table 2). The South Bosque Formation is composed pri-
Minor amounts of cross-laminated foraminiferal marily of medium-dark–to-medium gray silty mu-
packstone–grainstone, nodular foraminiferal packstone– drocks and includes beds of massive foraminiferal
grainstone, and massive argillaceous mudrock are calcareous mudrock and laminated calcareous fora-
also found within this unnamed member miniferal lime mudstone (Table 2). This unit, which
(Table 2). This unit is poorly exposed in outcrop, forms the top of the Eagle Ford Group in the study
and the full succession is observed only in core, area, averages 16 ft (5 m) in thickness.

Fairbanks et al. 391


Figure 9. North–south litho-
stratigraphic cross section of the
study area. The vertical columns
are described cores and out-
crops with color shading repre-
senting facies continuity. Note
the degree of facies and thick-
ness variability within each
stratigraphic unit. Location of
wells and outcrops is provided in
Figure 1. ACC = ACC 1 core;
Fm = formation; Mbr = member.

Cyclicity and Facies Continuity to 10 mi (16 km) (Figure 9). Correlation between
cores was conducted based on facies characteristics
This study uses a unique data set that provides an and associations, as defined previously in the Facies
excellent insight into variations in lateral facies con- and Stratigraphy sections.
tinuity. The 10 cores and outcrops in the Austin study High degrees of facies discontinuity are observed
area (Figure 1) constitute an approximately 11 mi, within each stratigraphic interval, even in close spac-
north–south transect. The measured sections from ings (50 ft [15 m]). For example, a nodular fora-
these cores and outcrops allow facies continuity to be miniferal packstone–grainstone bed in core 514 at 75
evaluated at several scales, ranging from 50 ft (15 m) ft (23 m), or 14 ft (4 m) above the top of the Buda, is

392 Eagle Ford Stratigraphy and Facies Architecture


not present 50 ft (15 m) to the northwest in core 500 trace metal, Mo, has been widely used as a proxy for
(Figure 9). Likewise, a massive bentonitic claystone benthonic redox potential because of its generally
bed at 79 ft (24 m), or 14 ft (4 m) above the top of the strong enrichment in organic-rich marine facies de-
Buda, in core 500 is not present in core 514 (Figure 9). posited under oxygen-depleted, sulfide-rich conditions
Additionally, two repetitive beds of massive argilla- (Algeo and Lyons, 2006; Tribovillard et al., 2006; Algeo
ceous mudrock observed in core 204 at 109 ft (33 m), and Tribovillard, 2009). In anoxic marine systems, Mo
or 7 ft (2 m) above the top of the Buda, are not shows significant variations from normal oceanic con-
represented 500 ft (150 m) to the south in core 302 ditions. For example, normal open-oceanic conditions
(Figure 9). contain 80%–100% of the Mo concentration in the
Facies continuity decreases substantially with Saanich Inlet of British Colombia, Canada; 70%–80%
distance. For example, 73% of units can be success- in the Framvaren Fjord of Norway; and just 3%–5%
fully correlated across a distance of 500 ft (152 m), in the Black Sea (Algeo and Rowe, 2012).
35% are traceable across 1 mi (1.6 km), and only Other redox-sensitive trace elements include U,
16% of beds are correlatable across 10 mi (16 km) V, Mn, and chromium (Cr), which some have linked
(Figure 9). At spacings of 10 mi (16 km) and greater, to changing oxygen conditions during the Ceno-
bed-scale correlations based on inorganic geochemical manian and Turonian (Tribovillard et al., 2006; Smith
data (XRD, XRF), organic geochemical data (Rock- and Malicse, 2010; Algeo and Rowe, 2012). In this
Eval TOC), and well-log data (GR) are suspect. It is study, Mo, U, V, Mn, and Cr were all assessed as
possible to correlate and define the boundaries of the geochemical paleoredox proxies. Of these trace ele-
four Eagle Ford stratigraphic units (Pepper Shale, un- ments, Mo is selected as a representative proxy for
named unit, Bouldin Member, and South Bosque anoxic bottom water conditions.
Formation) with reasonable confidence through the Major excursions and enrichments in these key
study area (Figure 9). trace elements occur in the Bouldin Member
Cyclicity of facies is readily observed within the (Figure 10), suggesting a period of maximum bottom
Eagle Ford, with the highest degrees of alternating water anoxia during this time. The Mo values, for
units contained within the Bouldin Member of the example, reached 50 ppm within the Bouldin
Lake Waco Formation. Beds ranging in thickness Member, compared with the typical 5–10 ppm
from 1 in. (2.5 cm) to 1 ft (30 cm) alternate between throughout the other members (Figure 10). Another
cross-laminated foraminiferal packstone–grainstone small enrichment in the Mo curve is observed
and laminated foraminiferal wackestone, with local within the upper part of the unnamed unit, suggest-
massive bentonitic claystones (Figure 9; Table 2). ing another interval of anoxia (Figure 10).
Several individual cycles can be laterally traced at
scales less than 1 mi (1.6 km), like those between the
Walnut Creek outcrop and the ACC core at ap- Regional Facies Architecture: Regional
proximately 95–99 ft (29–30 m), or 26–30 ft (8–9 m) Lithostratigraphic Correlation
above the top of the Buda (Figure 9).
The four-component stratigraphy described in the
previous section comprises the major depositional
Chemostratigraphic Analysis successions of the Eagle Ford Group in Central
Texas (Figure 4). In the deeper subsurface where
Energy-dispersive XRF analysis provided elemental these members have not been defined, the Eagle
composition data that were used to evaluate the in- Ford is typically divided into informal upper and
organic geochemistry of the Eagle Ford system. Var- lower units. However, it is possible to correlate the
iations in elemental composition have implications for Central Texas stratigraphy into the deeper sub-
ocean water chemistry as well as depositional setting. surface by comparing facies defined in the current
Certain trace metals are used as proxies for bottom study to those in the San Marcos arch area defined
water anoxia because they are redox sensitive and by Harbor (2011) (see Table 1). This comparison
relatively immobile in the sediment, thus preserving (Figure 11) illustrates that the Pepper Shale, the
primary signals (Algeo and Rowe, 2012). One such Lake Waco Formation (unnamed unit and Bouldin

Fairbanks et al. 393


Figure 10. Composite log and
type section for the ACC 1 core
showing the borehole gamma
ray (GR) log compared with the
computed gamma ray (CGR)
(potassium-thorium) from x-ray
fluorescence (XRF) data, the
identified facies from core de-
scription, a calcium curve from
XRF data suggesting mineralogy,
a molybdenum curve from XRF
data as a paleoredox proxy, and a
percent total organic carbon (%
TOC) curve as measured from
core. Note that the greatest pa-
leoredox enrichment occurs
within the Bouldin Member of
the Lake Waco Formation, which
is associated with elevated Ca
and high energy facies, and not
with greatest enrichment of %
TOC, which might be expected in
the unnamed unit.

Members), and the South Bosque Formation can all grainstone and laminated foraminiferal wackestone
be defined regionally in cored wells in the subsurface. relative to lower-energy mudrocks. Southward, that
Three cores, Hendershot, Orts, and Schauer, are used ratio diminishes, such that finer, low-energy mudrock
in conjunction with the ACC core in Austin, Texas, to facies become more abundant in the Schauer core
provide a north–south transect of nearly 70 mi (Figure 11).
(128 km) (Figure 11). South Bosque lithologies remain fairly un-
All stratigraphic intervals experience a south- changed from Austin southward. The Schauer core
ward downdip thickening along the axis of the arch. (Figure 11) reveals that facies of the South Bosque,
The Pepper Shale thickens from 5 ft (1.5 m) in Austin massive (and laminated) argillaceous foraminiferal
to approximately 12 ft (3.6 m) in the Schauer core mudrock (which are also facies of the unnamed
(Figure 11), at which point it undergoes a facies unit), become the dominant facies of the Eagle
change from massive argillaceous mudstone to lami- Ford Group.
nated calcareous foraminiferal lime mudstone toward The general thickening of individual units and
the shelf margin (Figures 2, 11). The unnamed unit the entire Eagle Ford Group toward the south is pro-
also displays thickening southward along the arch, bably associated with differential accommodation
from 10 ft (3 m) near Austin to 16 ft (5 m) in the around the San Marcos arch. Distal locations likely
Schauer core (Figure 11). experienced greater subsidence than the arch did
The Bouldin Member not only thickens south- (Figure 11). The facies change observed in the Bouldin
ward from approximately 13 ft (4 m) (ACC core) Member unit from high planktonic sediment con-
to approximately 25 ft (7.5 m) (Schauer core) centration facies in Austin to the argillaceous and
(Figure 11), but also it displays a general facies change. foraminiferal mudrock facies toward the Schauer core
In the Austin study area, there is a higher ratio of (Figure 11) is interpreted to result from a change in
high-energy cross-laminated foraminiferal packstone– planktonic productivity. Oxic zone productivity likely

394 Eagle Ford Stratigraphy and Facies Architecture


Figure 11. Cross section AA9
based on lithostratigraphy. This
subsurface correlation is based
on facies observed in core. The
C. J. Hendershot 1, H. P. Orts 2,
and F. T. Schauer et al. 1 cores
were described in detail by
Harbor (2011). The section
shows a general continuity of
stratigraphic intervals along the
San Marcos arch and thickening
southward. Location of transect
is provided in Figure 1. The up-
permost unit observed by Har-
bor (2011) refers to a transitional
Eagle Ford–Austin Chalk facies.
Modified from Harbor (2011).
ACC = ACC 1 core.

decreased to the south, based on the relative lack of Use of Gamma Ray Logs for Correlation
planktonic debris in the Schauer core.
Harbor (2011) described a transitional Eagle Core–Log Relationships in Study Area
Ford–Austin Chalk interval that is not represented in Because of the lack of conventional borehole GR logs
Austin but overlies the South Bosque Formation in for the cores in the study, pseudo GR logs were
the subsurface (Figure 11). This unit contains dis- constructed from elemental abundances of Th, U,
rupted bedded foraminiferal packstone and adjacent and K gathered from XRF. Both CGR, which com-
cross-laminated foraminiferal packstone–grainstone bines the Th and K responses, and total GR, which
(Harbor, 2011) (Figure 11). The lack of this unit in combines the Th, K, and U responses, were generated
the Austin study area suggests the presence of an (Table 3). Recent studies have shown that spectral
unconformity at the top of the Eagle Ford, marked GR logs are critical for the calibration and application
by a period of erosion. Alternatively, this unit of GR logs to identifying mineralogy and facies in
observed by Harbor (2011) could also represent a mudrocks (Rowe and Ruppel, 2013). Because of the
facies change from massive foraminiferal calcareous abundant K in the minerals, CGR logs are good in-
mudrock in the ACC core to disrupted bedded fora- dicators of clay mineralogy abundance in mudrocks.
miniferal packstone and cross-laminated foramini- However, GR logs provide a more complex indica-
feral packstone–grainstone in the Hendershot and tion of rock properties, because they combine the
Orts cores. mineral response of K in clay minerals with an

Fairbanks et al. 395


Table 3. Data for the ACC 1 Core, Containing Depth of Sample, Potassium Percentage, Uranium Parts per Million, and Thorium Parts per
Million from X-Ray Fluorescence, As Well As Measured Percent Total Organic Carbon

Depth (ft) Depth (m) K (%) U (ppm) Th (ppm) TOC (%) Depth (ft) Depth (m) K (%) U (ppm) Th (ppm) TOC (%)

70.4 21.46 0.545 -12 3 n/a 97.5 29.72 0.665 3 4 1.48


71.1 21.67 0.469 -12 2 n/a 98.4 29.99 0.148 -18 1 2.48
71.5 21.79 0.417 -13 2 n/a 98.6 30.05 1.407 7 1 0.24
72.0 21.95 1.548 -3 3 n/a 99.4 30.30 0.331 -11 1 3.29
72.4 22.07 1.372 0 4 n/a 99.8 30.42 1.278 9 4 0.28
72.9 22.22 1.389 1 5 n/a 100.3 30.57 1.460 -3 4 n/a
73.4 22.37 0.433 -14 3 n/a 101.2 30.85 0.195 -14 3 0.71
74.1 22.59 0.583 -12 3 n/a 101.8 31.03 1.118 -1 3 4.63
74.7 22.77 0.542 -11 3 0.22 102.6 31.27 1.210 1 3 4.56
75.5 23.01 0.382 -16 2 n/a 103.5 31.55 1.245 1 3 4.51
75.8 23.10 0.318 -11 2 n/a 103.8 31.64 0.147 -12 1 0.58
76.3 23.26 0.380 -9 4 0.15 104.7 31.91 1.267 -4 4 4.07
77.0 23.47 0.783 -8 3 n/a 105.5 32.16 1.229 0 2 5.40
77.7 23.68 0.673 -12 3 n/a 105.8 32.25 0.718 7 2 n/a
78.3 23.87 0.852 -4 4 0.60 106.2 32.37 1.715 -1 2 n/a
78.8 24.02 0.520 -14 2 n/a 106.5 32.46 1.318 1 4 6.48
79.7 24.29 1.091 -5 4 n/a 106.9 32.58 0.280 -11 1 1.11
80.1 24.41 1.538 0 3 n/a 107.4 32.74 1.336 -1 3 3.89
80.5 24.54 1.866 4 5 2.13 107.9 32.89 0.225 -14 3 n/a
80.9 24.66 1.834 0 3 n/a 108.8 33.16 0.564 -6 4 0.32
81.4 24.81 2.063 2 5 n/a 109.4 33.35 1.466 11 3 6.47
81.9 24.96 2.133 3 5 n/a 110.1 33.56 1.599 11 7 n/a
82.4 25.12 1.612 1 5 1.34 110.9 33.80 0.969 3 2 8.91
82.9 25.27 1.593 1 6 n/a 111.5 33.99 1.400 2 6 7.16
83.5 25.45 1.711 7 6 n/a 112.4 34.26 1.549 9 5 5.37
84.1 25.63 1.522 0 5 n/a 113.0 34.44 0.234 -14 2 1.20
84.5 25.76 1.420 0 5 2.81 113.7 34.66 1.847 9 5 7.99
85.6 26.09 1.627 -2 4 n/a 114.3 34.84 1.398 6 4 0.88
86.1 26.24 1.695 -1 4 n/a 115.5 35.20 0.981 -10 5 6.29
86.6 26.40 1.589 -7 6 3.16 116.5 35.51 1.440 16 6 5.20
87.5 26.67 1.650 4 4 n/a 117.3 35.75 1.512 7 6 5.92
88.6 27.01 1.336 2 3 3.75 118.7 36.18 1.449 9 6 1.00
89.5 27.28 1.266 -2 3 n/a 119.5 36.42 1.449 15 13 1.31
90.4 27.55 1.405 0 6 3.62 120.5 36.73 1.901 15 13 1.20
91.5 27.89 1.141 2 5 n/a 121.3 36.97 1.813 7 15 1.20
92.4 28.16 1.675 9 5 2.05 122.5 37.34 1.887 7 15 1.47
92.8 28.29 1.614 -2 4 n/a 123.6 37.67 1.308 9 11 3.96
93.4 28.47 1.803 2 4 n/a 124.3 37.89 1.212 14 11 2.33
94.5 28.80 1.762 1 5 4.34 124.8 38.04 1.740 12 15 n/a
95.5 29.11 0.433 -13 1 n/a 125.2 38.16 0.355 -13 3 0.20
96.3 29.35 1.163 0 4 n/a 126.3 38.50 0.098 -20 2 0.13
96.6 29.44 1.268 2 4 0.31 127.7 38.92 0.069 -14 0 0.22

Abbreviations: K = potassium; n/a = no data available; Th = thorium; TOC = total organic carbon; U = uranium.

396 Eagle Ford Stratigraphy and Facies Architecture


indication of bottom water redox conditions from foraminiferal calcareous mudrock and laminated calca-
the U response. reous foraminiferal lime mudstone. The CGR values for
Examination of the Central Texas Eagle Ford both of these units are higher than those of the Bouldin,
Group illustrates differences in spectral GR response indicating higher clay mineral abundance (Figure 10).
that can be related to the four-component stratig-
raphy defined in the ACC 1 core (Figure 10). The Regional Log-Based Correlations
basal Pepper Shale is typically defined by a high CGR Comparison of GR log trends in the ACC cored well
and GR response, especially compared with the with other subsurface wells extending to the south
underlying Buda Limestone (Figure 10). The high across the San Marcos arch suggests that general
CGR and GR values exhibited by the Pepper Shale correlations of facies may be possible (Figure 12).
are a function of the massive (and laminated) cal- However, the GR logs available for these wells do
careous foraminiferal lime mudstone facies, which not closely follow core-based lithological relationships
contain abundant illite, kaolinite, and smectite (Figure 12). Based on these observations, it must be
(Figure 10). concluded that facies determination founded solely on
The overlying unnamed unit is characterized by total GR logs (K–Th–U) is liable to be misleading. As
lower CGR and GR responses caused by higher illustrated in Figure 10, a moderate total GR response
carbonate and lower clay mineral abundances, in the ACC core corresponding to the unnamed unit
compared with the Pepper Shale (Figure 10). The that is followed by a high GR zone corresponding
similarity between the values of the CGR and the to the Bouldin Member could be interpreted as a
GR indicates that little U is present in the massive- carbonate-rich interval succeeded by a clay mineral–
to-laminated calcareous foraminiferal lime mud- rich interval. However, core, CGR, and total GR
stone of the unnamed unit. response reveal the opposite facies relationships
The Bouldin Member is unique among the Eagle (Figure 10). Thus, GR logs alone are insufficient for
Ford units in that the CGR and GR logs from its rocks facies determination in the Eagle Ford, and CGR logs
display a significant separation (Figure 10). The low must be used for accurate facies recognition.
CGR values indicate low clay mineral content. This Because sequence stratigraphic interpretations
is consistent with the higher carbonate content in hinge on accurate determination of facies and accom-
the high-energy, highly calcareous, laminated fora- modation, correlations and facies designations based
miniferal wackestone–packstone and cross-laminated solely on total GR logs are likely to produce erroneous
foraminiferal packstone–grainstone facies that con- results. For example, although the high total GR of the
stitute this unit. The high frequency oscillations in Bouldin Member appears generally correlative, what, if
both CGR and GR curves reflect the cyclic alter- anything, can be reliably interpreted about sediment
nations in these two facies. The separation of the supply, water depth, accommodation, or even miner-
CGR and GR in the Bouldin identifies an abundance alogy from the total GR log? Without knowledge of
of U in these rocks (Figure 10), which is reflected in the concentration of and variations in U in the unit
the high total GR. The implications of the high U (made possible by comparison of CGR and total GR
concentrations are considerable when trying to use logs), none of these key factors that are necessary for
total GR logs to interpret or correlate facies. Note stratigraphic correlation (using sequence concepts or
that, based solely on a total GR log, most workers not) can be reliably defined.
would interpret the very high GR response in this
interval as an indication that the rocks are higher in
clay mineral content than any other part of the Eagle Total Organic Carbon Analysis
Ford. Instead, the Bouldin contains the highest car-
bonate and lowest clay mineral concentrations of the The Upper Cretaceous Eagle Ford is a proven source
entire Eagle Ford (Figure 10). rock in Texas (Robison, 1997). Samples collected in
Like the unnamed unit, the South Bosque For- the current study for source rock analysis vary in
mation displays similar low values for both the CGR TOC stratigraphically (Figure 10; Table 3) but con-
and GR, suggesting low U content (Figure 10). Like tain an average of 2.4% (range 0.1%–8.4%) TOC. At
the unnamed unit, the dominant facies are massive the base of the Eagle Ford, a dramatic increase in

Fairbanks et al. 397


Figure 12. Cross section BB9
based on gamma ray (GR) and
pseudo computed GR (CGR) log
(ACC core). The GR log sig-
natures are somewhat recog-
nizable to the southwest across
the San Marcos arch. ACC = ACC
1 core; Fm = formation; Mbr =
member.

TOC is observed in comparison with the underlying basin environment and ideal conditions for organic
Buda Limestone (Figure 10). The Pepper Shale dis- matter preservation (Demaison and Moore, 1980).
plays a general upward increase in TOC, with an Recognition criteria for restricted basin conditions in-
average of 2.9% (range 1.7%–6.3%) TOC (Figure 10). clude the presence of black, organic rich shales con-
The overlying unnamed unit displays a large range in taining high TOC; a lack of benthonic fauna; and
organic enrichment but contains the highest overall enrichment in redox-sensitive trace elements. However,
average of 4.0% (range 0.1%–8.4%) TOC (Figure 10). the correspondence of these characteristics is not ob-
In contrast, the Bouldin Formation is characterized by served in the Central Texas Eagle Ford succession,
the lowest TOC, with an average of only 2.3% (range substantiating the recent findings that paleoceano-
0.4%–4.1%) TOC (Figure 10), a finding that is cor- graphic changes controlling anoxia were more complex
roborated by Corbett and Watkins (2013), who ob- than previously suggested (Corbett and Watkins, 2013).
served sharp declines in TOC within the Bouldin. The Maximum TOC enrichment is observed in the
overlying South Bosque Formation displays an aver- unnamed unit (average TOC 4%, maximum 8.4%)
age increase to 2.4% (range 0.2%–5.4%) TOC and (Figure 10). Facies within the unnamed unit domi-
then gradually decreases upward into the Austin nantly consist of massive foraminiferal calcareous
Chalk (Figure 10). mudrock and laminated calcareous foraminiferal
lime mudstone (Figure 10). The association between
relatively low-energy facies and high TOC would
DISCUSSION suggest that this interval represents the most distal
setting and the greatest degree of basin restriction.
According to conventional wisdom in mudrock suc- However, geochemical proxies for anoxia (enrichment
cessions, an association exists between a restricted in Mo, U, Mn, V, Cr) indicate that greatest bottom

398 Eagle Ford Stratigraphy and Facies Architecture


water anoxia occurred during deposition of the when compared with the Bouldin Member. Recent
Bouldin Member (Figure 10). Surprisingly, the investigators (Ratcliffe et al., 2012) suggested that
Bouldin Member is composed of the highest energy TOC can be estimated from paleoredox proxies,
grain-rich facies (Figures 7, 10). These grain-rich facies particularly enrichment in Mo. However, in the
are not traditionally expected to form within restricted current study, the disparity between paleoredox
basin conditions. The Bouldin Member is also the in- indicators and low TOC in the Bouldin Member
terval characterized by the highest total GR and highest suggests that TOC estimation from inorganic chem-
U concentrations but the lowest TOC (Figure 10). istry data may provide misleading results.
Hence, two unlikely associations are observed: (1) high High-energy facies within the Bouldin Mem-
GR signatures with corresponding calcareous high- ber are recognized in core and outcrop by cross-
energy facies, and (2) lowest TOC values corre- laminations, dune foreset laminations, low angle
sponding to maximum basin restriction (Figure 10). ripple laminations, planar-to-subplanar laminations,
Planktonic globigerinid foraminifera are the mud rip-up clasts, pebble and bioclastic lags, and
dominant grain type within the Bouldin Member. scour and truncation surfaces. Facies variability is also
This sediment type was sourced from the photic zone most pronounced in the Bouldin Member, even at
and reflects surface water environmental conditions, small scales (50 ft [15 m]). Bedding thickness var-
not bottom water conditions. The high concentration iations, pinchouts, erosional scouring and trunca-
of planktonic debris indicates accentuated pro- tions, and localized nodular facies are all recognized
ductivity during this interval, providing high rates of as facies variability in the Bouldin Member. These
carbonate sediment supply. As mentioned earlier, a current-induced sedimentary structures and facies
strong association exists between the massive bentonitic variabilities do not reflect shallow water processes
claystone layers and calcareous foraminiferal units. In but are interpreted to result from bottom-current
fact, the massive bentonitic claystones (interpreted to reworking. Bottom currents, which are capable of
result from volcanic ash settling into marine waters) are transporting grains up to fine sand and even gravels
generally restricted to the Bouldin Member. This close (Faugéres and Stow, 1993; Stow et al., 2002;
relationship suggests that heightened productivity in Frébourg et al., 2016), can be independent of
the oxic zone of the water column is stimulated by eustatic control. Furthermore, many of the minor
nutrient enrichment from volcanic input, and it drives components, consisting of benthonic fauna (mainly
sediment supply independent of eustatic forcing. inoceramids), live in dysoxic environments and can be
Notwithstanding paleoredox conditions suggest- transported into more anoxic settings, as supported
ing maximum bottom water anoxia and ideal con- by the fragmented nature of the grains.
ditions for organic preservation (Figure 10), TOC Although extremely calcareous and represented
enrichment is lowest in the Bouldin Member, aver- by the highest energy facies, the Bouldin Member
aging 2.3% compared with a 4.0% average in the is interpreted to have been deposited in a marine
unnamed unit (Figure 10). The discrepancy between setting with substantial bottom water anoxia. This
high preservation potential and low TOC during interpretation contrasts with classical sequence strati-
deposition of the Bouldin Member is interpreted graphic perspectives that interpret such high-energy,
to result from carbonate dilution. Frequent influx, carbonate-rich deposits as forming proximal to an
higher associated sedimentation rates, and relatively active carbonate platform. Expected characteristics of
rapid deposition of foraminiferal tests resulted in a deposit from such a shallow water setting would be
grain-rich carbonate facies within which very little increased carbonate sediment supply, high-energy fa-
TOC was preserved. The high proportion of carbo- cies from storm waves or turbidity currents, and de-
nate beds to mudrock beds within the Bouldin creased TOC resulting from sediment dilution and
Member resulted in low TOC values overall for this poor preservation in an oxygenated environment. The
interval. Additionally, higher calcite content from Bouldin Member displays these characteristics. How-
the planktonic material caused the Bouldin Member ever, the presence of current-induced sedimentary
to become more resistant to postdepositional com- structures from deep water bottom-current reworking
paction. Greater compaction in other Eagle Ford (Frébourg et al., 2016), high degrees of facies vari-
units allowed for differentially concentrated TOC ability, and paleoredox proxies demonstrate that this

Fairbanks et al. 399


traditional model does not apply to the Central Texas water nutrient enrichment, stimulated in part by
Bouldin Member. volcanic activity. Bottom-current activity results from
Defining mineralogies and the facies they rep- the interplay of thermohaline and wind-driven cur-
resent is not simple, because over-reliance on total rents because of both Coriolis forces and seafloor
GR logs (K–Th–U) can provide very misleading topography (Stow et al., 2002; Frébourg et al., 2016).
concepts of mineralogy and facies distribution. As These two primary controls are processes that may
observed in Figure 10, the unnamed unit and South be independent from eustatic fluctuation, rendering
Bosque Formation exhibit moderate total GR values, classical sequence stratigraphic applications unreli-
whereas the intervening Bouldin Member displays able in the Eagle Ford system of Central Texas.
much higher values. Based solely on these GR pat-
terns, the following interpretations could justifiably
but erroneously be drawn: (1) the unnamed unit and SUMMARY AND CONCLUSIONS
South Bosque successions contain the highest levels
of carbonate in the Eagle Ford and should contain The Upper Cretaceous (Cenomanian–Turonian)
low levels of TOC, and (2) the Bouldin Member Eagle Ford Group of Central Texas contains the
contains lower levels of carbonate and higher following seven distinct facies: (1) massive argillaceous
amounts of clay minerals, suggesting that these rocks mudrock, (2) massive foraminiferal calcareous mu-
accumulated in a more distal, low-energy setting, drock, (3) laminated calcareous foraminiferal lime
where high levels of TOC would be expected. Core mudstone, (4) laminated foraminiferal wackestone, (5)
data show the relationships to be just the opposite. cross-laminated foraminiferal packstone–grainstone,
Even when mineralogy is well constrained (e.g., (6) massive bentonitic claystone, and (7) nodular
by spectral GR logs), the conditions under which foraminiferal packstone–grainstone. These seven facies
these facies were deposited cannot be readily defined comprise four stratigraphic packages defined herein as
from logs. Conventional thinking regarding many the Pepper Shale, the unnamed unit of the Lake Waco
carbonate-rich successions is that carbonate is sourced Formation, the Bouldin Member of the Lake Waco
from shallow water settings and that high levels Formation, and the South Bosque Formation.
of carbonate indicate relatively shallow water sed- The basal Pepper Shale is an argillaceous, fine-
imentation regardless of the setting. Abundant car- grained (clay-sized) unit. The unnamed unit, newly
bonate in the Eagle Ford, however, is the result of designated in this study, is an argillaceous and
two processes not interpreted to be connected with calcareous (foraminiferal), massive mudrock. The
water depth: nutrient driven productivity and bottom- Bouldin Member is a high-energy, carbonate-rich
current reworking. Accordingly, attempts to place (foraminiferal) interval with high degrees of facies
Eagle Ford rocks into a simple sequence stratigraphic variability. The uppermost unit is the South Bosque
framework based on an assumed connection between Formation, which, like the unnamed unit, is an
mineralogy (i.e., carbonate content) and accom- argillaceous and calcareous (foraminiferal) massive
modation or water depth are of questionable merit. and laminated mudrock.
Alone, GR profiles are insufficient in the Eagle The Bouldin Member is defined by high API
Ford for determination of (1) facies, (2) TOC content, total GR values and low API CGR (GR K–Th)
(3) depositional environment, and (4) sequence strati- values. The disparity between total GR and CGR
graphic correlations. Comparison of the GR curves with values is commensurate with U content. Enrichment
CGR curves provides Eagle Ford investigators with in U, as well as in other paleoceanographic proxies
more accurate means for rock character determination. (i.e., Mo, V, Cr), suggests that the Bouldin Member
Primary controls on Eagle Ford character are represents maximum basin restriction, despite being
(1) sediment supply from planktonic foraminifera composed of the most carbonate-rich, highest energy
and (2) depositional reworking by bottom-current facies. The comparison of total GR logs to CGR logs
activity. Globigerinid foraminifera comprise the is requisite, because GR alone may provide mis-
dominant carbonate sediment type and originate in leading determination of facies, TOC content, dep-
the photic zone of the water column. Changes in ositional environment, and sequence stratigraphic
carbonate sedimentation rates are driven by surface implications.

400 Eagle Ford Stratigraphy and Facies Architecture


High energy deposition within the Bouldin Mem- Algeo, T. J., and T. W. Lyons, 2006, Mo–total organic carbon
ber resulted from bottom-current activity. Carbonate covariation in modern anoxic marine environments:
Implications for analysis of paleoredox and paleohydro-
content was controlled by heightened productivity graphic conditions: Paleoceanography, v. 21, doi:
in the oxic zone, which led to a higher sediment 10.1029/2004PA001112.
supply of planktonic foraminiferal skeletal debris. Algeo, T. J., and H. Rowe, 2012, Paleoceanographic ap-
Both bottom-current reworking and planktonic plications of trace-metal concentration data: Chem-
sediment supply were controlling factors that were ical Geology, v. 324–325, p. 6–18, doi:10.1016
/j.chemgeo.2011.09.002.
decoupled from eustatic sea-level fluctuations, thus Algeo, T. J., and N. Tribovillard, 2009, Environmental analysis of
rendering classical sequence stratigraphy unreliable paleoceanographic systems based on molybdenum-uranium
in the Eagle Ford system. covariation: Chemical Geology, v. 268, p. 211–225, doi:
Even at a small lateral spacing (50 ft [15 m]), 10.1016/j.chemgeo.2009.09.001.
much variability of Eagle Ford facies is observed in Barnes, V. E., 1992, Geologic map of Texas: Austin, Texas,
Bureau of Economic Geology, scale 1:500,000, 4 sheets,
cores and outcrops, and this is attributed to bottom- SM0003.
current reworking and planktonic productivity. Childs, O. E., G. Steele, and A. Salvador, 1988, Gulf Coast
Bottom-current reworking is responsible for erosional region, correlation of stratigraphic units of North America
scouring, truncation, and localized distribution of (COSUNA) project: AAPG, oversize chart, 20 map sheets.
facies. Planktonic productivity, possibly resulting Corbett, J. M., and D. K. Watkins, 2013, Calcareous nan-
nofossil paleoecology of the mid-Cretaceous Western
from the nutrient enrichment of volcanic ash settling Interior Seaway and evidence of oligotrphic surface waters
into marine waters, influenced sediment supply. during OAE2: Palaeogeography, Palaeoclimatology,
Facies continuity decreases substantially with Palaeoecology, v. 392, p. 510–523, doi:10.1016/j
distance. For example, 73% of units can be success- .palaeo.2013.10.007.
fully correlated across a distance of 500 ft (152 m), Culotta, R., T. Latham, M. Sydow, J. Oliver, L. Brown, and
S. Kaufman, 1992, Deep structure of the Texas Gulf
35% are traceable across 1 mi (1.6 km), and only 16% Passive Margin and its Ouachita-Precambrian basement:
of beds are correlatable across 10 mi (16 km). At Results of the COCORP San Marcos arch survey: AAPG
spacings of 10 mi (16 km) and greater, bed scale Bulletin, v. 76, no. 2, p. 270–283.
correlations based on inorganic geochemical data Dawson, W. C., 1997, Limestone microfacies and sequence
stratigraphy: Eagle Ford Group (Cenomanian-Turonian)
(XRD, XRF), organic geochemical data (Rock-Eval
North-Central Texas outcrops: Gulf Coast Association of
TOC), and well-log data (GR) are suspect. However, Geological Societies Transactions, v. 47, p. 99–105.
it is possible at these distances to define and correlate Dawson, W. C., B. J. Katz, L. M. Liro, and V. D. Robison,
the approximate boundaries of the four Eagle Ford 1993, Stratigraphic and geochemical variability: Eagle
stratigraphic units, at least in the study area. Ford Group, east-central Texas, in Rates of geologic pro-
cesses, tectonics, sedimentation, eustasy and climate—
Contrary to conventional schemas, mudrock dep-
Implications for hydrocarbon exploration: 14th Annual
osition is complex, involving the interplay of many Gulf Coast Section of the Society of Economic Pale-
controlling processes. Facies variability, often over- ontologists and Mineralogists Foundation Research
looked by stratigraphers and explorationists, is a signif- Conference, Houston, Texas, December 5–8, 1993,
icant aspect of the Eagle Ford system and has p. 19–28, doi:10.5724/gcs.93.14.0019.
Demaison, G. J., and G. T. Moore, 1980, Anoxic environ-
implications for source rock quality, seal capacity, res-
ments and oil source bed genesis: AAPG Bulletin, v. 64,
ervoir characterization, and hydraulic fracture potential. no. 8, p. 1179–1209.
Caution must be employed when evaluating mudrock Denne, R. A., and J. A. Breyer, in press, Integrated stratigraphy
systems, because nodular facies, ash beds, and other and biostratigraphy of the Cenomanian-Turonian interval
heterogeneities can dramatically influence correlatability. (Eagle Ford and Woodbine groups), Texas. Part 2: Re-
gional depositional episodes of the Cenomanian-Turonian
Eagle Ford and Woodbine groups of Texas, in J. A. Breyer,
ed., The Eagle Ford Shale—A renaissance in U.S. oil
production: AAPG Memoir 110.
REFERENCES CITED Denne, R. A., R. E. Hinote, J. A. Breyer, T. H. Kosanke,
J. A. Lees, N. Englelhardt-Moore, J. M. Spaw, and
Adkins, W. S., and F. E. Lozo, Jr., 1951, Stratigraphy of the N. Tur, 2014, The Cenomanian-Turonian Eagle Ford
Woodbine and Eagle Ford, Waco area, Texas: Southern Group of South Texas: Insights on timing and
Methodist University Fondren Science Series 4, paleoceanographic conditions from geochemisty
p. 105–161. and micropaleontologic analyses: Palaeogeography,

Fairbanks et al. 401


Palaeoclimatology, Palaeoecology, v. 413, p. 2–28, Hentz, T. F., W. A. Ambrose, and D. C. Smith, 2014, Ea-
doi:10.1016/j.palaeo.2014.05.029. glebine play of the southwestern East Texas basin:
Donovan, A. D., and S. T. Staerker, 2010, Sequence strat- Stratigraphic and depositional framework of the
igraphy of the Eagle Ford (Boquillas) Formation in the Upper Cretaceous (Cenomanian-Turonian) Woodbine
subsurface of South Texas and outcrops of West Texas: and Eagle Ford Groups: AAPG Bulletin, v. 98, no. 12,
Gulf Coast Association of Geological Societies Trans- p. 2551–2580, doi:10.1306/07071413232.
actions, v. 60, p. 861–899. Hentz, T. F., and S. C. Ruppel, 2010, Regional lithostratig-
Doveton, J. H., and D. F. Merriam, 2004, Borehole petro- raphy of the Eagle Ford Shale: Maverick Basin to East
physical chemostratigraphy of Pennsylvanian black shales Texas Basin: Gulf Coast Association of Geological So-
in the Kansas subsurface: Chemical Geology, v. 206, cieties Transactions, v. 60, p. 325–337.
p. 249–258. Jiang, M. J., 1989, Biostratigraphy and geochronology of the
Dravis, J. J., 1980, Sedimentology and diagenesis of the Upper Eagle Ford Shale, Austin Chalk, and Lower Taylor Marl in
Cretaceous Austin Chalk Formation, South Texas and Texas based on calcareous nannofossils, Ph.D. dissertation,
northern Mexico, Ph.D. thesis, Rice University, Houston, Texas A&M University, College Station, Texas, 524 p.
Texas, 513 p. Jones, M. T., and S. R. Gislason, 2008, Rapid releases of metal
Dubiel, R. F., P. D. Warwick, L. Biewick, L. Burke, salts and nutrients following the deposition of volcanic
J. L. Coleman, K. O. Dennen, and C. Doolan et al., 2010, ash into aqueous environments: Geochimica et Cosmo-
Geology and assessment of undiscovered oil and gas re- chimica Acta, v. 72, p. 3661–3680, doi:10.1016
sources in Mesozoic (Jurassic and Cretaceous) rocks of /j.gca.2008.05.030.
the onshore and state waters of the Gulf of Mexico region, Kearns, T. J., 2011, Chemostratigraphy of the Eagle Ford
U.S.A.: Gulf Coast Association of Geological Societies Formation, M.S. thesis, The University of Texas at Ar-
Transactions, v. 60, p. 207–216. lington, Arlington, Texas, 254 p.
Duggen, S., P. Croot, U. Schacht, and L. Hoffmann, 2007, Langmann, B., K. Zaksek, M. Hort, and S. Duggen, 2010,
Subduction zone volcanic ash can fertilize the surface Volcanic ash as fertilizer for the surface ocean: Atmos-
ocean and stimulate phytoplankton growth: Evidence pheric Chemistry and Physics, v. 10, p. 3891–3899, doi:
from biogeochemical experiments and satellite data: 10.5194/acp-10-3891-2010.
Geophysical Research Letters, v. 34, p. L01612–L01617, Liro, L. M., W. C. Dawson, B. J. Katz, and V. D. Robison,
doi:10.1029/2006GL027522. 1994, Sequence stratigraphic elements and geochemical
Duggen, S., N. Olgun, P. Croot, L. Hoffmann, H. Dietze, variability within a “condensed section”: Eagle Ford
P. Delmelle, and C. Teschner, 2010, The role of airborne Group, East-Central Texas: Gulf Coast Association of
volcanic ash for the surface ocean biogeochemical iron- Geological Societies Transactions, v. 44, p. 393–402.
cycle: A review: Biogeosciences, v. 7, p. 827–844, doi: Lundquist, J. J., 2000, Foraminiferal biostratigraphic and
10.5194/bg-7-827-2010. paleoceanographic analysis of the Eagle Ford, Austin, and
Fairbanks, M. D., 2012, High resolution stratigraphy and facies lower Taylor Groups (middle Cenomanian through
architecture of the Upper Cretaceous (Cenomanian- lower Campanian) of Central Texas, Ph.D. thesis, The
Turonian) Eagle Ford Group, Central Texas, M.S. thesis, University of Texas at Austin, Austin, Texas, 544 p.
The University of Texas at Austin, Austin, Texas, 120 p. Pessagno, E. A., 1969, Upper Cretaceous stratigraphy of the
Faugères, J.-C., and D. A. Stow, 1993, Bottom-current- western gulf coast area of Mexico, Texas, and Arkansas:
controlled sedimentation: A synthesis of the contourite Geological Society of America, v. 111, p. 59–64.
problem: Sedimentary Geology, v. 82, p. 287–297, doi: Phelps, R. M., 2011, Middle-Hauterivian to Lower-
10.1016/0037-0738(93)90127-Q. Campanian sequence stratigraphy and stable isotope
Feray, D. E., and K. Young, 1949, Guidebook, Cretaceous of geochemistry of the Comanche platform, South Texas,
Austin, Texas area, 1949: 17th annual field trip by Ph.D. dissertation, The University of Texas at Austin,
Shreveport Geological Society: Shreveport, Louisiana, Austin, Texas, 240 p.
Shreveport Geological Society, p. 51–56. Pierce, J. D., 2014, U-Pb geochronology of the Late Creta-
Frébourg, G., S. C. Ruppel, R. G. Loucks, and J. Lambert, ceous Eagle Ford shale, Texas; defining chronostrati-
2016, Depositional controls on sediment body archi- graphic boundaries and volcanic ash source, M.S. thesis,
tecture in the Eagle Ford/Boquillas system: Insights from The University of Texas at Austin, Austin, Texas, 144 p.
outcrops in west Texas, United States, AAPG Bulletin, Pratt, L. M., 1984, Influence of paleoenvironmental factors
doi:10.1306/12091515101. on preservation of organic matter in Middle Cretaceous
Galloway, W. E., 2008, Depositional evolution of the Gulf of Greenhorn Formation, Pueblo, Colorado: AAPG Bulle-
Mexico sedimentary basin, in A. D. Miall, ed., The tin, v. 68, no. 9, p. 1146–1159.
sedimentary basins of the United States and Canada: New Ratcliffe, K. T., A. M. Wright, and K. Schmidt, 2012, Ap-
York, Elsevier, p. 505–549, doi:10.1016/S1874-5997 plication of inorganic whole-rock geochemistry to shale
(08)00015-4. resource plays: An example from the Eagle Ford Shale
Harbor, R. L., 2011, Facies characterization and stratigraphic Formation, Texas: Sedimentary Record, v. 10, no. 2,
architecture of organic-rich mudrocks, Upper Cretaceous p. 4–9, doi:10.2110/sedred.2012.2.4.
Eagle Ford Formation, South Texas, M.S. thesis, The Robison, C. R., 1997, Hydrocarbon source rock variability
University of Texas at Austin, Austin, Texas, 184 p. within the Austin Chalk and Eagle Ford Shale (Upper

402 Eagle Ford Stratigraphy and Facies Architecture


Cretaceous), East Texas, U.S.A.: International Journal of 16, 2011, http://www.searchanddiscovery.com/pdfz
Coal Geology, v. 34, p. 287–305, doi:10.1016/S0166- /abstracts/pdf/2010/intl/abstracts/ndx_smith.pdf.html.
5162(97)00027-X. Sohl, N. F., E. Martı́nez, P. Salmercón-Urreña, and
Rowe, H., N. Hughes, and K. Robinson, 2012, The quantifi- F. Soto-Jaramillo, 1991, Upper Cretaceous, in A. Sal-
cation and application of handheld energy-dispersive vador, ed., The Gulf of Mexico Basin: The geology of
x-ray fluorescence (ED-XRF) in mudrock chemo- North America, v. J: Boulder, Colorado, Geological
stratigraphy and geochemistry: Chemical Geology, v. Society of America, p. 205–244.
324–325, p. 122–131, doi:10.1016/j.chemgeo.2011.12.023. Stow, D. A., J.-C. Faugeres, J. A. Howe, C. J. Pudsey, and
Rowe, H., and Ruppel, S. C., 2013, High-resolution chemo- A. R. Viana, 2002, Bottom currents, contourites and
stratigraphy of the Bakken Formation, Williston Basin (abs.): deep-sea sediment drifts: Current state-of-the-art, in
AAPG International Conference and Exhibition, Cartagena, D. A. Stow, C. J. Pudsey, J. A. Howe, J.-C. Faugères,
Colombia, September 8–11, 2013, accessed June 9, 2014, and A. R. Viana, eds., Deep-water contourite systems:
http://www.searchanddiscovery.com/abstracts/html/2013 Modern drifts and ancient series, seismic and sedimentary
/90166ice/abstracts/rowe.htm. characteristics: Geological Society, London, Memoirs
Ruppel, S. C., R. G. Loucks, and G. Frébourg, 2012, Guide to 2002, v. 22, p. 7–20.
field exposures of the Eagle Ford-equivalent Boquillas Tribovillard, N., T. J. Algeo, T. Lyons, and A. Riboulleau,
Formation and related Upper Cretaceous units in 2006, Trace metals as paleoredox and paleoproductivity
southwest Texas: The University of Texas at Austin, proxies: An update: Chemical Geology, v. 232, p. 12–32,
Bureau of Economic Geology, Mudrock Systems Re- doi:10.1016/j.chemgeo.2006.02.012.
search Laboratory Field-Trip Guidebook, 151 p. Tyler, N., and W. A. Ambrose, 1986, Depositional systems
Salvador, A., and J. M. Q. Muñeton, 1989, Stratigraphic and oil and gas plays in the Cretaceous Olmos Formation,
correlation chart of Gulf of Mexico Basin, in A. Salvador, South Texas: Report of Investigations 152: Austin, Texas,
ed., The Gulf of Mexico Basin: The geology of North Bureau of Economic Geology, University of Texas at
America, v. J: Boulder, Colorado, Geological Society of Austin, 42 p.
America, plate 5. Young, K., 1977, Guidebook to the geology of Travis County:
Scott, R. W., 2010, Cretaceous stratigraphy, depositional Austin, Texas, The University of Texas at Austin, ac-
systems, and reservoir facies of the northern Gulf of cessed October 6, 2011, http://www.lib.utexas.edu/geo
Mexico: Gulf Coast Association of Geological Societies /ggtc/toc.html.
Transactions, v. 60, p. 597–609. Young, K., 1986, Cretaceous marine inundations of
Smith, C. N., and A. Malicse, 2010, Rapid handheld x-ray the San Marcos Platform, Texas: Cretaceous Re-
fluorescence (HHXRF) analysis of gas shales: AAPG search, v. 7, p. 117–140, doi:10.1016/0195-6671(86)
Search and Discovery article 90108, accessed March 90013-3.

Fairbanks et al. 403

Anda mungkin juga menyukai