ArticlePDF Available

Lakes, Loess, and Paleosols In the Permian Wellington Formation of Oklahoma, U.S.A.: Implications For Paleoclimate and Paleogeography of the Midcontinent

Authors:

Abstract and Figures

Lower to mid-Permian deposits of the Midcontinent (U.S.A.) record a significant and long recognized aridification because they archive the shift from more humid facies (e.g., coal, organic shale) of the Pennsylvanian to widespread redbeds, semiarid to seasonal paleosols (Calcisols, Vertisols), and evaporites by the mid-Permian. The provenance, transport and depositional processes of the voluminous Permian redbeds of the Midcontinent, however, remain largely undefined. The Artinskian Wellington Formation in Oklahoma exhibits high-frequency cycles comprising organic-rich laminated mudstone with thinly laminated (inferred primary) dolomite, variegated laminated mudstone with gypsum, massive, red to gray-green mudstone with pedogenic overprinting, and pale red siltstone. The gypsum exhibits a distinct 87/86Sr isotopic ratio (0.709199) that is inconsistent with Permian seawater. We suggest that these facies record deposition in ephemeral to perennial lakes during a time of increasing aridity and seasonality, the latter indicated by abundant mudcracks, vertic-type paleosols, conchostracans, and lungfish burrows. The fine and uniform grain size and the geochemistry of the siliciclastic component suggest far-travelled and likely eolian transport that ultimately accumulated in both subaqueous and subaerial environments. Provenance analysis indicates the siliciclastic component was sourced chiefly from the southeastern Ouachita– Appalachian orogen and the Ancestral Rocky Mountains (ARM) or its derivative sediment.
Content may be subject to copyright.
Journal of Sedimentary Research, 2013, v. 83, 825–846
Research Article
DOI: 10.2110/jsr.2013.59
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA, U.S.A.:
IMPLICATIONS FOR PALEOCLIMATE AND PALEOGEOGRAPHY OF THE MIDCONTINENT
JESSICA M. GILES,
1
MICHAEL J. SOREGHAN,
2
KATHLEEN C. BENISON,
3
GERILYN S. SOREGHAN,
2
AND
STEPHEN T. HASIOTIS
4
1
Chesapeake Energy Corporation, Oklahoma City, Oklahoma, U.S.A.
2
School of Geology and Geophysics, University of Oklahoma, Norman, Oklahoma, U.S.A.
3
Department of Geology and Geography, West Virginia University, Morgantown, West Virginia, U.S.A.
4
Department of Geology, University of Kansas, Lawrence, Kansas, U.S.A.
A
BSTRACT
:Lowertomid-PermiandepositsoftheMidcontinent(U.S.A.)recordasignificantandlongrecognized
aridification because they archive the shift from more humid facies (e.g., coal, organic shale) of the Pennsylvanian to
widespread redbeds, semiarid to seasonal paleosols (Calcisols, Vertisols), and evaporites by the mid-Permian. The provenance,
transport and depositional processes of the voluminous Permian redbeds of the Midcontinent, however, remain largely
undefined. The Artinskian Wellington Formation in Oklahoma exhibits high-frequency cycles comprising organic-rich
laminated mudstone with thinly laminated (inferred primary) dolomite, variegated laminated mudstone with gypsum, massive,
red to gray-green mudstone with pedogenic overprinting, and pale red siltstone. The gypsum exhibits a distinct
87/86
Sr isotopic
ratio (0.709199) that is inconsistent with Permian seawater. We suggest that these facies record deposition in ephemeral to
perennial lakes during a time of increasing aridity and seasonality, the latter indicated by abundant mudcracks, vertic-type
paleosols, conchostracans, and lungfish burrows. The fine and uniform grain size and the geochemistry of the siliciclastic
component suggest far-travelled and likely eolian transport that ultimately accumulated in both subaqueous and subaerial
environments. Provenance analysis indicates the siliciclastic component was sourced chiefly from the southeastern Ouachita–
Appalachian orogen and the Ancestral Rocky Mountains (ARM) or its derivative sediment.
INTRODUCTION
The late Paleozoic records Earth’s ‘‘best known’’ pre-Quaternary ice
age characterized by an ‘‘extreme’’ climate (Kutzbach and Gallimore
1989; Crowell 1999; Fielding et al. 2008). Attendant with Pangean
assembly, the equatorial Central Pangean Mountains formed and
atmospheric circulation shifted from zonal to monsoonal, resulting in
cross-equatorial flow and marked seasonality (Kutzbach and Gallimore
1989; Parrish 1993). For reasons that remain debated (Tabor and Poulsen
2008), a significant shift in climate from more humid to arid conditions
occurred in the Permian, especially in western equatorial Pangaea. In
Midcontinent North America, this aridity shift corresponds to the
deposition of the first regionally extensive redbed units by mid-Permian
time. The highly weathered nature of these mudstone and evaporite
outcrops, however, has hindered attempts to constrain the depositional
setting, provenance, and climatic implications (Chaplin 2004).
This study focuses on detailed analysis of a whole core of the
Wellington Formation (Fig. 1A) from northern Oklahoma (Fig. 1B),
with additional observations from a laterally extensive quarried outcrop
(Kay County Quarry) of the same interval. Our objective is to address the
depositional environments, source(s) and transport mechanism(s) of the
siliciclastic sediment and, thereby, paleoclimatic implications of this
regionally extensive unit. Specifically, this study assesses whether 1) the
facies reflect principally continental (lacustrine, loess, and paleosol)
deposition, rather than the previously interpreted marginal marine
environment; and 2) eolian transport was the chief mechanism for the
(volumetrically predominant) siliciclastic component of the Wellington
Formation. These findings bear on improving constraints on regional
paleoclimatic conditions, including atmospheric circulation in this part of
western Pangaea, during a time of major climatic transition in the late
Paleozoic.
GEOLOGIC SETTING AND CONTROVERSIES
The late Paleozoic Gondwanan–Laurasian plate collision produced the
equatorial Central Pangean Mountains (CPM; Scotese et al. 1979; Blakey
2007), expressed in southeastern–mid North America as the Appala-
chian–Ouachita foldbelt (Fig. 2). This collision was roughly coeval with
formation of the intracratonic Ancestral Rocky Mountains (ARM) in
western Pangaea (Fig. 2; Kluth and Coney 1981). Between these tectonic
domains, the Midcontinent region of the U.S. (central Pangaea) generally
formed a low-gradient, low-elevation landscape in nearly equatorial
latitudes (,5uSto,5uN; Tabor et al. 2008) during Pennsylvanian
through Permian time. Superimposed on this low-gradient paleogeogra-
phy was the Amarillo–Wichita uplift and adjacent Anadarko basin and
(northward) Hugoton embayment to the west and southwest and the
Ozark uplift and Ouachita uplifts to the east and southeast (Fig. 2;
McKee and Oriel 1967; Johnson et al. 1989). By Permian time, however,
several uplifts began to subside (De Voto 1980; Gilbert 1992; Sweet and
Soreghan 2010; Soreghan et al. 2012), such that the Amarillo and Wichita
uplifts were covered by lower–middle Permian strata (Johnson 1989; Price
et al. 1996; Gilbert 2002; Soreghan et al. 2012). Similarly, Upper
Pennsylvanian and Permian strata, including the Wellington Formation
in parts of Oklahoma, blanket the Precambrian-cored Nemaha ridge
Published Online: September 2013
Copyright
E
2013, SEPM (Society for Sedimentary Geology) 1527-1404/13/083-825/$03.00
(Fig. 2; Luza 1978; Chaplin 1988). This region remained equatorial (,5u
S to ,2uN) from Virgilian (Gzhelian) through Leonardian (Kungarian)
time (Tabor et al. 2008).
Through much of the late Paleozoic, ‘‘icehouse’’ conditions character-
ized by repeated waxing and waning of continental ice sheets (Veevers
and Powell 1987; Crowell 1999; Fielding et al. 2008; Fischbein et al. 2009)
produced pervasive cyclothems at low latitudes (Wanless and Shepard
1936; Heckel 1986; Veevers and Powell 1987). Upper Paleozoic strata of
the Midcontinent record both glacioeustasy and the onset of increasing
aridity (Tabor et al. 2008). From the Pennsylvanian to the mid-late
Permian, deposition in the Midcontinent shifted from limestone–shale
cyclothems of the Chase and Council Grove groups to redbeds and
evaporites of the Nippewalla and El Reno groups (Fig. 1; Johnson et al.
1989; Golonka and Ford 2000; Heckel 2007). The Leonardian
(Artinskian) Sumner Group, including the Wellington Formation
described here, marks the transition between the underlying limestone–
shale cyclothems and the overlying redbeds and evaporites.
The Wellington Formation of the Sumner Group crops out in a narrow
strip from northern Oklahoma into mid-Kansas (Fig. 1B; Mazzullo 1999;
Hall 2004). The Wellington Formation is 250 m (820 ft) thick (Raasch
1946) and encompasses six members; this study focuses on the Anhydrite
sequence and the Otoe and Midco members (Fig. 1A). Gypsum or
anhydrite, siltstone, shale, and mudstone with subordinate dolomudstone
and dolomite constitute the lithology of the Wellington Formation in
Oklahoma (Fig. 3; Raasch 1946; Boardman 1999; Chaplin 2004). The
Wellington and coeval deposits are commonly capped by the coarser-
grained Garber Sandstone (Fig. 1A) in Oklahoma, whereas northward
(Kansas) evaporites, including bedded halite in the Hutchinson Salt
(Fig. 1A; Fig. 2; Swineford and Runnels 1953; Jones 1965; Watney et al.
1988), overlie the Wellington.
Controversy exists over the depositional environment of the Wellington
Formation, with older studies suggesting deposition partly in a lacustrine
setting (Dunbar 1924; Raasch 1946; Tasch 1961a, 1964) and newer studies
postulating marine or marginal marine conditions (Mazzullo 1999;
Chaplin 2004; Hall 2004). Most studies inferring a marine origin for
the Oklahoma strata focused on the dolomite and evaporite facies, owing
to poor outcrop exposures that precluded detailed petrographic
descriptions of other facies. Chaplin (2004) provided general interpreta-
tions of the Wellington Formation based on newly obtained whole core
(including the core used in this study); however, he focused primarily on
the underlying Chase and Council Grove formations of unequivocal
marine origin. For the Wellington Formation, studies favoring a
F
IG
. 1.—A) Permian stratigraphy of outcrops in north-central Oklahoma, combined and modified from Dunbar et al. (1960), Wilson (1962), Chaplin (1988), and
Johnson et al. (2001). The units and lithology of the Wellington Formation and the Nolans Limestone are shown for Oklahoma and Kansas, based on the work of Ver
Wiebe (1937), Raasch (1946), O’Connor et al. (1968), and Chaplin (1988). The superscripts 1 and 2 in the inset column refer to the specific source that defined the units in
Kansas and Oklahoma. *Stratigraphic names of Wellington and Nolans are members unless otherwise stated, viz. Anhydrite and Basal sequences. B) Study area of Kay
County, Oklahoma depicting the location of the KC-1 core and Kay County Shale Pit outcrop (modified from Chaplin 2004).
826 J.M. GILES ET AL.
JSR
continental origin are based on selected paleontological aspects (e.g.,
insect beds) constrained to certain horizons. Few studies have focused on
the siliciclastic strata, which dominate by volume. Finally, recent fluid-
inclusion and sedimentology work by Benison and Goldstein (2001, 2002)
suggests that the overlying Nippewalla Group in Kansas consists of acid
saline lake deposits amid mudflat, eolian, and paleosol siliciclastic
deposits. A continental origin for even parts of the Wellington Formation
would significantly alter paleogeographic and paleoclimatic reconstruc-
tions and attendant depositional and transport pathways for this time and
region.
METHODS
This study centers on the KC-1 core from Kay County, Oklahoma
(Fig. 1B), drilled by the Oklahoma Geological Survey using freshwater as
the drilling fluid. The cored interval begins at the surface and extends
,91 m, penetrating the upper Anhydrite sequence, the Otoe Member,
and most of the Midco Member of the Wellington Formation, which
crops out at the surface (Fig. 3); thus, the Billings Pool and Antelope
Flats members were not present in the core. We slabbed intervals of the
core and described the entire core at centimeter-scale resolution.
Representative samples of all facies were selected for thin-section and
geochemical analysis. Additionalfaciesdataarebasedonfield
observations of the Wellington Formation exposed in the Kay County
Shale Pit (36u35953.310N, 97u2190.850W) immediately adjacent to the
core site (Chaplin 2004). This quarry exposes the Midco Member of the
Wellington in three dimensions over ,200 m. This aided in more
definitive descriptions of facies contacts and lateral variability of facies.
Four siliciclastic facies and two chemical sediment facies were identified
using macroscopic sedimentologic features supplemented by petrographic
and SEM observations and geochemical and mineralogical characteriza-
tion. Thirty-five thin sections were examined. The core description,
including facies thicknesses and vertical transitions, was used to calculate
facies abundances and determine the most common facies transitions
using a transition-count matrix, an approach similar to Markov-chain
analysis (Carr 1982; Driese and Dott 1984). Five representative samples
were disaggregated using sodium phosphate (Na
6
O
18
P
6
)andsonication
for 10 minutes for grain-size analysis. Although grain-size analysis on
lithified rocks can be problematic, the Permian rocks in north central
Oklahoma have not been buried very deeply, ,1km(Carteretal.1998).
The mudstones are poorly indurated, even in core, and petrographic
examination indicates that the facies are poorly cemented, such that the
sonicated sediment provides a reasonable approximation of the original
sediment texture. The measured grain size likely represents a maximum,
in the event any weak cements persisted. Analyses of the disaggregated
sediment after treatment was performed using a Beckman-Coulter LS-230
laser particle size analyzer. In addition, the mass and grain size of the
‘‘acid-insoluble-residue’’ (siliciclastic fraction) was determined from eight
dolomite samples through dissolution in 2N HCL to infer detrital influx
during carbonate deposition. Source-rock-analyzer-based TOC and
pyrolysis data were collected on 10 dark mudstone samples by Weath-
erford Laboratories.
F
IG
. 2.—Leonardian paleogeography of the Midcontinent showing main features mentioned in text. Arrows show general directions of zonal winds (seasonally shifting
from southeasterlies to northeasterlies as intertropical convergence zone migrates) and inferred seasonal monsoonal westerlies. Locations of geochemical datasets from
other Permian strata used in this study are indicated on map by diagonal lines. Locations and relative sizes of uplifts and basins are based on Berg (1977), McKee and
Oriel (1967), Johnson et al. (1989), Algeo and Heckel (2008), and Soreghan et al. (2012).
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 827
JSR
F
IG
. 3.—Stratigraphic column of the Wellington Formation in the KC-1 core illustrating facies, inferred water level, and the boundary of the members encompassed by
the cored interval.
828 J.M. GILES ET AL.
JSR
T
ABLE
1.— Summary descriptions and interpretations of the main facies in the Permian Wellington Formation.
Facies Grain Size Lithology Color Features Mineralogy Environmental Interpretation
Siliciclastic
Facies
Siltstone
Lithofacies
fine silt 10
m
m
(st dev 5.6)
Siltstone Light brown to
pale orange
Massive to faintly
laminated or chaotic
bedding +/2patchy
fabric, rare root traces
Primarily subangular
to angular quartz;
rutile, muscovite,
zircon, chlorite, K-
feldspar, and pyrite
Subaerial (eolian)
suspension deposition;
overprinting by
pedogenesis, disruption
by salt crusts
Shallowing-
upward ‘‘cycle’’
Massive
Mudstone
Lithofacies
Red massive
mudstone
very fine silt
to clay
4
m
m
(st dev 5.2)
Mudstone;
gypsum
Dark red to
purple
Massive bedding +/2
slickensides & laminae
of gypsum; local
ped-like structures
Primarily clay (illite
and smectite)
subangular to
angular quartz;
chlorite, rutile,
K-feldspar
Both subaerial (eolian
and minor unconfined
flow deposits) and
subaqueous suspension
deposition. Extensive
pedogenesis.
Gray massive
mudstone
very fine silt
to clay
4
m
m
(st dev 5.2)
Lt gray and/or
greenish or
bluish gray
Primarily clay (illite
and smectite)
angular quartz,
muscovite; quartz,
chlorite, rutile,
muscovite, and
biotite
Laminated
Mudstone
Lithofacies
Variegated
mudstone
clay 3
m
m
(st dev 5.3)
Mudstone;
dolomite;
gypsum
Variable and
mixed: red, grey,
yellow, brown,
white
Laminated (mm-scale)
to wavy or chaotic
bedding with evaporite
and dolomite
laminae +/2small
slickensides, mudcracks,
burrows, root traces,
brecciated fabrics, rare
conchocostracans, rare
cubic casts
Primarily clay (illite
and smectite) some
angular quartz
grains; albite, iron
oxides, quartz,
muscovite, chlorite,
and pyrite
Low-energy, subaqueous
suspension settling into
perennial lake, likely
shallow, some
desiccation and
pedogenesis
Organic-rich
mudstone
very fine silt
to clay
4
m
m
(st dev 5.2)
Mudstone;
dolomite
Lt gray - black;
some yellow
and/or green
Laminated (mm-scale)
dolomite and dark
mudstone
Mudstone: Primarily
clay (illite and
smectite) Few
angular quartz grains;
muscovite and
pyrite. Dolomitic
micrite
Low-energy subaqueous
suspension settling in
perennial-lake
environment. Likely
deeper with less
evidence for
desiccation
Chemical
Facies
Dolomite
Lithofacies
n/a ,1–5
m
m Dolomite;
Ca/Mg 5,1–
1.2; ,25%
siliciclastics
White - buff Laminated (mm-scale)
or thicker beds +/2
desiccation cracks, rare
burrows,
microcrystalline texture,
fenestral fabric
Accessory
mineralogy 5
zircon, quartz,
K-feldspar, albite,
chlorite, iron oxides,
celestite, barite,
pyrite
Low-energy subaqueous
perennial-lake
environment.
Gypsum
Lithofacies
n/a n/a Gypsum;
relatively pure
Clear, white, or
stained red
Laminated (small-scale)
or thicker beds
intermixed with
mudstone +/2vertically
oriented crystals +/2
displacive crystals in
mudstone.
Gypsum and/or
anhydrite
Subaqueous deposition
(bottom growth and
laminated habits) to
penecontemporaneous
formation (displacive
habits); High-salinity
water
______________________
R
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 829
JSR
830 J.M. GILES ET AL.
JSR
Qualitative phase identification on representative samples from each
facies was accomplished by backscattered electron imaging coupled with
energy-dispersive X-ray analysis (EDXA) using 20 kV accelerating
voltage and 10 nA beam current (see details in Pack 2010). In addition,
35 samples from siliciclastic facies throughout the core were powdered,
processed to eliminate carbonate-bound Ca (following Soreghan and
Soreghan 2007), and analyzed for major and trace elements using
combined XRF and ICP-MS. Finally, a Rigaku automated wide-angle X-
ray diffractometer (XRD) with a graphite monochromator and copper
tube (CuK, 40 kV, 30 mV) was used to characterize clay mineralogy; runs
were conducted between 4 and 70u2hwith step size of 0.05u2hand 5 s
dwell time. Oriented aggregate mounts for XRD were prepared following
Poppe et al. (2002).
For detrital-zircon geochronological analysis, we used a siltstone
sample from the middle of the Midco Member of the Wellington collected
from the quarry outcrop (to obtain sufficient sample from a specific
stratum). For analysis, 2 kg of the siltstone was processed using standard
crushing and heavy-mineral separation techniques (Gehrels et al. 2008).
The analytical procedures in employing the Micromass Isoprobe at the
University of Arizona closely followed those described in detail by
Dickinson and Gehrels (2008) except that a spot diameter of 25
m
mwas
utilized because of the fine grain size of the zircon. A crystal of the Sri
Lankan zircon (564 64Ma(2herror)) was analyzed for calibration
every five unknowns. The interpreted age for grains ,1 Ga are taken
from
206
Pb/
238
Uageswhereasgrains.1 Ga are based on
206
Pb/
207
Pb
ages; these ages are summarized in Appendix 1 (see Acknowledgments).
Age uncertainties are reported at the 1-sigma level and include only
measurement or analytical errors. Ages were analyzed for discordance by
comparing the
206
Pb/
207
Pb ages to the
206
Pb/
238
Uagesforthosegrains
.1 Ga; a 30% difference was used to filter the data. In total, 79 grains
were retained for analysis, and the data are presented as a cumulative
probability plot in which the age and associated errors of each grain are
summed.
For strontium isotope analyses, gypsum samples were selected
representing pristine (unrecrystallized) examples of the various morphol-
ogies present (laminated, displacive, and bottom-growth crystals).
External surfaces were cleaned and the gypsum was powdered using a
drill equipped with an ultra-fine bit. Approximately 2 mg of each sample
powder was placed in 4 ml of quartz-distilled water for 1 week until no
visible sample remained. Subsequently, 0.8 ml of Seastar 14 molar NHO
3
was added to each vial, to make the solution approximately 3 N HNO
3
.
The samples were slowly pumped through 50 microliter Sr spec columns
equilibrated with 3 N HNO
3
. Approximately 1 ml more 3 N HNO
3
was
used to wash the other ions from the columns. The Sr was eluted with 2 ml
of quartz-distilled water and dried on a hotplate with 1 drop of dilute
phosphoric acid. The samples were loaded on Re center filaments with a
Ta activator and run at approximately 1400uCinmulti-dynamicmodeon
aIsotopXPhoeniX64.Asmallcorrectionof0.000006wasappliedtoeach
sample to account for the slight difference between the measured value of
NBS987 and the recommended value of 0.710248.
RESULTS
Results are presented in two parts: 1) description and interpretation of
facies comprising the KC-1 core (summarized in Table 1), supplemented
with outcrop observations; 2) geochemical and detrital-zircon geochro-
nologic characterization of the Wellington siliciclastic facies for inferring
provenance.
Facies Analysis
Dolomite
Description.—The dolomite mudstone facies is very fine grained (clay to
very fine silt), white to light gray, and occurs as (1) microlaminae
intercalated with siliciclastic facies (Fig. 4A1, Fig. 5A), or (2) thicker (4 to
48 cm, average 17 cm), structureless to vaguely laminated beds (Fig. 4A2)
that compose 2% of the core. Insoluble residue of the dolomite mudstone
averages 25.2% (n55), and consists predominantly of quartz, potassium
feldspar, iron oxides, and pyrite. Microprobe images reveal aggregates of
rhombohedral dolomite crystals (,1–5
m
m) with a Ca/Mg ratio of ,1–
1.2. Secondary minerals include pyrite and Fe oxides in addition to trace
amounts of various silicate minerals. Common thin vertical cracks and/or
rare infilled vertical structures occur (Fig. 4B). Gypsum locally occurs as
thin interbeds in the dolomite (Fig. 4B). In outcrop, the dolomite is laterally
continuous (Fig. 6A), with sharp upper and lower bases (Fig. 6B, C).
Interpretation.—The dolomite is interpreted as primary, because it (1)
occurs as discrete beds, particularly as laminae interlayered with
siliciclastic mudstone, (2) displays a very fine-grained (micritic) texture
(Last 1990), and (3) lacks any petrographic evidence for replacement
textures or dolomitization of biogenic calcite or aragonite (Garcia del
Cura et al. 2001). Primary dolomite is typically microcrystalline and
characterized by a fine (less than about 10
m
m) grain size (Folk 1973; von
der Borch 1976; Vasconcelos and McKenzie 1997). Dolomite with a
larger grain size and a saccharoidal texture is commonly secondary and
displays evidence of replacement (von der Borch 1976). Primary dolomite
of lacustrine origin is typically very fine grained (,1–2
m
m) and forms
aggregates (Last 1990), like the Wellington dolomite. The coexistence of
iron oxides and pyrite in the dolomite and associated laminated facies is
consistent with dolomite formation from metabolic activity of bacteria
(Trudinger et al. 1985; Last 1990; Machel 2004; Sanz-Montero et al.
2009). In low temperatures and anoxic conditions, bacterial sulfate
reduction is the main process that produces pyrite (Trudinger et al. 1985;
Williams-Stroud 1994; Machel 2004). During bacterial sulfate reduction,
metabolic CO
2
and bicarbonate ions are released, providing ideal
conditions for carbonate formation in general (Sanz-Montero et al.
2009) and primary dolomite in particular (Vasconcelos and McKenzie
1997; Sanchez-Roman et al. 2008, 2009).
The intercalation of millimeter- to submillimeter-scale laminae of
dolomite and siliciclastic material suggests subaqueous suspension
deposition into a quiet-water environment. There are no indications of
flaser, ripple, or any other type of stratification indicative of traction
processes, and individual beds are laterally continuous across the quarry
face. There are, however, indications of exposure or shallow, subaqueous
conditions superimposed on the dolomite mudstone facies. For example,
we interpret the very fine vertical cracks (Fig. 4A2) as mudcracks,
indicating that exposure and periodic wetting and drying occurred after
the dolomite laminae formed. The larger infilled vertical structures
(Fig. 4B) are likely vertical burrows that suggest oxic (shallower-water)
conditions existed subsequent to dolomite formation.
r
F
IG
.4.CharacteristicsedimentaryfeaturesofvariousfaciesintheWellingtoncore.A1) ,1-mm-thick dolomite laminae (227.4 ft, 69.3 m; Otoe), and A2) thicker,
bedded dolomite (white or light) ,1.25 cm thick with desiccation cracks (51.0 ft, 15.5 m; Midco). B) Microcrystalline dolomite with evaporite(?) voids, laminations of
gypsum, and a burrow (22.5 ft, 6.9 m; Midco). C) Gypsum with vertical crystal growth. C1) Aswallow-tailgypsumcrystal(248.0ft,75.6m;Otoe);C2) Amuddrapeover
gypsum crystals; both are indicative of subaqueous, primary origin (61.1 ft, 18.6 m; Midco). D) Thin section of vertically growing crystals capped by dissolution surface,
suggesting subaqueous origin (60.0 ft, 18.3 m; Midco). E) Photomicrograph displaying well-sorted angular to subangular quartz grains of siltstone facies (279.3 ft, 85.1 m;
Otoe). F) Mound-shaped structure within the siltstone facies (128.0 ft, 39 m; Otoe).
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 831
JSR
832 J.M. GILES ET AL.
JSR
Pure dolomite of stoichiometric composition has a Ca/Mg ratio of 1;
therefore, the Ca/Mg ratio of the dolomite here suggests a relatively
stoichiometric dolomite (Chilingar 1957; Last 1990; Machel 2004) similar
to modern dolomite in saline lakes, although the stoichiometric
composition of dolomite remains equivocal as the chief means of
distinguishing between marine and lacustrine environments (Last 1990).
Gypsum
Description.—Gypsum occurs in three forms (in order of abundance),
1) thicker beds intercalated with siliciclastic material that display
vertically oriented crystals, 2) laminae in mudstone facies (discussed
below), and 3) displacive nodular gypsum in mudstone. The thicker
gypsum beds compose 3% of the Wellington Group in the core. The
vertically oriented crystals widen upward, originate from a common
plane, and exhibit a sharp contact with the overlying mudstone (Fig. 4C1,
D). The overlying mudstone thickens over depressions and thins over
raised gypsum crystal terminations (Fig. 4C2). The transition to overlying
mudstone typically consists of a layer of gypsum regrowth over a thin
layer of mud or a sharp contact with truncated crystals. Rare nodular
beds of gypsum occur with vague remains of vertically twinned gypsum
rhombs. No replacement textures have been identified in thin-section or
microprobe analysis. Microprobe analysis indicates nearly pure calcium
sulfate, with less than 0.1 wt% other (mostly silicate) phases. Strontium
isotope data (
87
Sr/
86
Sr) collected from five gypsum samples, representing
all three occurrences of gypsum and lacking any textural evidence of
replacement, are listed in Table 2. The values range from 0.708688 to
0.709711 and average 0.709199.
Interpretation.—The vertically aligned (swallow-tail) crystals of gyp-
sum record subaqueous bottom precipitation in saline surface water
(Benison et al. 2007). The upward-widening habit of the vertical gypsum
crystals and their origination from a common plane support a primary,
subaqueous origin as bottom-growth crystals (Gibert et al. 2007). The
thickening and thinning of mudstone directly above the vertical gypsum
crystals indicates that subaqueous gypsum precipitation occurred prior to
deposition of the mud, which infilled microtopography (Spencer and
Lowenstein 1990). Sharp contacts with the overlying mudstone that
truncate the tops of the underlying gypsum crystals suggest dissolution
prior to mud deposition.
The displacive gypsum crystals likely reflect penecontemporaneous
formation in the host mud. Similar displacive habits form in the shallow
subsurface in both modern saline lakes and sabkhas (Spencer and
Lowenstein 1990; Benison et al. 2007). Some of the displacive gypsum
beds, however, do appear to have altered from anhydrite to gypsum (or
vice versa), which readily occurs at shallow burial depths (Warren 1989;
Gibert et al. 2007).
The Sr-isotope data from all gypsum morphologies (laminated,
displacive, and bottom growth) are inconsistent with (and significantly
higher than) Permian seawater values reported by Korte et al. (2006),
indicating that the gypsum did not precipitate from waters directly
connected to marine sources. This is discussed in more detail below, but
the data indicate a significant continental source for the Sr, and thus,
strong evidence for a continental origin rather than a supratidal or
sabkha setting for gypsum formation.
Siltstone
Description.—The very fine siltstone facies varies from pale orange to
brown to locally white to light gray, and composes 8% of the core. It
consists of well-sorted, angular to subangular siliciclastic (mostly quartz)
grains (Fig. 4E) of fine silt (10
m
m); the clay-size fraction ranges from
,19 to 70% (average ,34%). The siltstone facies occurs in massive beds
10 cm to 1.2 m thick, with common soft-sediment deformation and
chaotic bedding (Fig. 4F). Downwardly bifurcating, vertical traces (2 to
5cmlong)occurlocally.Typically,basalcontactsareslightlygradational
with the underlying massive mudstone facies, whereas upper contacts are
typically sharp and capped by the organic or variegated mudstone facies.
Although the siltstone facies is present through the middle and lower
portion of the core, it is absent in the upper ,33 m. In the quarry
outcrop, the siltstone appears massive and highly fractured, and is
laterally continuous, with mottling, peds, and vertical traces along the
upper surface.
Interpretation.—The fine-grained and consistently well-sorted texture
suggests eolian transport for this facies (Pye 1987; Tsoar and Pye 1987).
The very fine grain size reflects longer-term suspension and, therefore,
transport from a relatively distant source (Pye 1987; Smalley et al. 2005).
In addition, massive bedding and gradational contacts are most
consistent with suspension fallout during eolian transport as opposed
to traction deposition (Pye 1987; Johnson 1989). Rare root traces (i.e.,
rhizoliths) and poorly developed pedogenic overprinting (see below)
record exposure and preclude subaqueous deposition. The soft-sediment
deformation in this facies could reflect seismites (Montenat et al. 2007)
whereas the chaotic bedding may be a result of extensive bioturbation
(Smoot and Lowenstein 1991). An alternate method of forming soft-
sediment deformation, chaotic bedding, and patchy fabrics, however, is
through precipitation of efflorescent salt crusts. These salt deposits are
typically very thin, easily dissolved, and rarely preserved (Smoot and
Castens-Seidell 1994). In highly permeable sediment, such as silt, the
growth of efflorescent salt results in mounds ranging from 2 to 4 cm that
deform subjacent structures and create a ‘‘popcorn like’’ texture that can
trap silt in depressions, as seen in both modern saline pans (Smoot and
Castens-Seidell 1994) and Permian saline deposits (Benison and Goldstein
2000, 2001).
Organic Mudstone Facies
Description.—The organic mudstone facies is predominantly dark gray
to greenish black (Fig. 5B, 6B) and composes 23% of the study section.
The average grain size is very fine silt to clay (4
m
m). Light-colored, thin
laminae of subangular to angular coarse silt quartz grains occur
commonly. Total organic carbon (TOC) values are 0.78–1.09 wt.%,
averaging ,0.92 wt%, and Rock-Eval data indicate immature, type IV
kerogen. Microprobe (Fig. 5A) and cathodoluminescence imaging
indicate (sub-millimeter- to millimeter-scale) planar laminae consisting
of alternating dark-colored calcareous mudstone and light-colored
dolomite, locally convoluted. Muscovite, pyrite, and quartz occur in the
clastic laminae, similar to the mineralogy of the gray massive mudstone.
XRD analysis indicates less quartz compared to the red-colored facies
and a clay-rich matrix, consisting of micas, kaolinite, and chlorite.
r
F
IG
. 5.—A) SEM-BSE image of dolomite laminae in organic mudstone facies (dolomite produces brighter reflectance in CL; 219.4 ft, 66.9 m; Otoe). B) Typical laminae
in organic mudstone facies exhibiting color variations of gray, black, and lighter-colored dolomite. Notice possible oil staining (24.9 ft, 7.6 m; Midco). C) Typical
expression of laminae in the variegated mudstone facies. Notice color variation among laminae (144.8 ft, 44.1 m; Otoe). D) Cubic halite casts (11.1 ft, 3.4 m; Midco) in
variegated mudstone facies. E) Wavy to sinuous complex mudcracks in variegated mudstone facies viewed in cross section (239.5 ft, 73.0 m; Otoe). F) Imprints of
conchostracans, sizes range from ,1–5 mm (179.0 ft, 54.6 m; Otoe). G) Photomicrograph showing ‘‘floating’’ quartz grains in clay matrix (153.4 ft, 46.8 m; Otoe) of
massive mudstone facies. H) Slickensides (205 ft, 62.5 m; Otoe) in massive mudstone facies.
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 833
JSR
F
IG
.6.OutcropphotosoffaciesinMidcoMember.A) Photo across quarry face illustrating the thin-bedded, continuous nature of the facies and lack of scouring. B)
Organic mudstone facies sharply overlying a thicker dolomite bed. C) Interbedded dolomite and organic mudstone. D) Outcrop of massive mudstone showing gradational
nature at base from underlying variegated mudstone facies. E) Blocky peds and reduction features in the massive mudstone facies.
834 J.M. GILES ET AL.
JSR
Dolomite laminae are abundant in this facies, occurring as either
laminae, beds up to 8 cm thick, or small round clasts (Figs. 5A, B, 6C). In
contrast to the abundant mudcracks in the laminated mudstone facies
(described below), this facies rarely exhibits vertical cracks. Overlying
facies are mainly the laminated mudstone or locally the massive mudstone
or siltstone facies. Basal contacts are typically sharp from the underlying
massive mudstone or siltstone facies or rarely the variegated mudstone.
Interpretation.—The dark color, elevated TOC values, and planar
lamination indicate a low-energy, low-oxygen, subaqueous environment.
Although such conditions are commonly associated with deep water, they
can also occur in shallow basins if salinity gradients are sufficient to
reduce vertical mixing (e.g., Great Salt Lake, Utah; Hardie et al. 1978).
The convolute laminae could reflect seismic activity (Montenat et al.
2007), or perhaps mass-flow processes (Benison and Goldstein 2001),
although the regional, low-relief setting argues against the slopes
necessary for the latter. The alternating high- and low-TOC mudstone
laminae suggest seasonality during deposition (Allen and Collinson 1986),
whereas the dolomite laminae represent increased carbonate input (Last
1990). As noted above, the fine-scale, relatively pure laminae of dolomite
interlaminated with the organic mudstone suggests subaqueous, suspen-
sion settling.
Laminated Mudstone Facies
Description.—Laminated, variegated mudstone is the most common
facies in the core, composing 30% of the section. Individual beds of very
fine silt to clay (3–4
m
m) laminae average ,50 cm thick. Various colors,
chiefly red with less common orange, white, brown, yellow, and gray
occur, varying among millimeter scale laminae (Fig. 5C). Planar to wavy
laminae at the sub-millimeter to millimeter scale are most common, but
chaotic and/or brecciated intervals also occur. Lamination typically fines
upward and consists of light-colored very fine silt, dolomite, or red or
gray mud; rare laminae of peloids also occur. Dolomite appears in the
form of small-scale (,0.1–6 cm) planar laminae, small rounded clasts, or
disseminated within the mudstone laminae. Gypsum is common in this
facies and occurs as planar beds 1–2 cm thick, or as displacive crystals.
Rare cube-shaped casts occur on bedding planes in the upper 3.5 m of
core (Fig. 5D).
Rare, downwardly bifurcating traces occur in this facies along with thin
vertical cracks (1–3 mm wide), randomly oriented slickensides, and
chaotic bedding (Fig. 5E). Rare imprints of whole branchiopod
crustacean conchostraca (,1–5 mm in diameter; identified by Dr. P.
Murry, Tarleton State University, personal communication, 2010; Dr. R.
Scott, University of Tulsa, personal communication, 2010) occur on
bedding planes (Fig. 5F); these conchostraca are whole and unfragmen-
ted and exhibit well-developed growth lines. Generally, the organic or
(rarely) massive mudstone facies underlies this facies, whereas the massive
mudstone or siltstone facies gradationally overlie this facies.
Interpretation.—The millimeter scale laminae and conchostracans
indicate deposition in a low-energy subaqueous environment (Tasch
1964; Smoot 1993). In the Permian, conchostracans typically lived in
continental environments and tolerated a range of salinities (Tasch 1964;
Swanson and Carlson 2002; Park and Gierlowski-Kordesch 2007; Scott
2010). The taphonomy (unabraded and unfragmented) of these bioclasts
indicates little transport prior to death. Tasch (1961a, 1961b, 1962) and
Swanson and Carlson (2002) noted that conchostracans are common in
the Wellington Formation on bedding planes of dolomite–mudstone
intervals.
The laminae of dolomite are similar in texture and grain size to the
thicker laminated (inferred primary) dolomite. The presence of very fine-
grained, laminated dolomite supports subaqueous deposition of the
laminated mudstone facies. Also, the thin, planar laminae of gypsum with
vertically oriented crystals are likely primary, further supporting a
subaqueous origin. The cubic casts record halite casts (Benison and
Goldstein 2001); the alignment of these cubes along bedding planes
suggests that the halite formed in the water column and settled to the
bottom.
Although the evidence suggests a predominance of subaqueous
deposition, periodic subaerial exposure is indicated by 1) root traces
and slickensides, which reflect pedogenesis (Retallack 1990), and 2)
mudcracks and brecciated fabrics, which further indicate subaerial
exposure. The irregular (wavy) shape of the mudcracks is consistent with
highly saline environments (Hardie et al. 1978), whereas the brecciated
fabric could indicate incipient pedogenesis.
Massive Mudstone Facies
Description.—The massive mudstone facies is variably dark red to
purple or light gray to greenish with reddish to purple hues more common
in the Otoe Member and drab gray colors more common in the younger
Midco Member (Fig. 6D). Regardless of color, all display massive bedding
with well-developed, randomly oriented slickensides and blocky peds
(Fig. 6E), rare laminae of gypsum, local brecciated fabrics, and local,
floating subangular to angular quartz grains within the muddy matrix.
Dark red to pale purple mudstone composes 19% of the core and
generally forms beds about 0.5 m thick. Gray to greenish massive
mudstone forms beds about 35 cm thick on average and composes 16% of
the study section. The massive mudstone is composed of very fine silt to
clay grains (3–4
m
m) of mica and clay minerals (illite, kaolinite, and
smectite) and rare dolomite. Petrographic observations also reveal a clay-
rich matrix with local outsized silt grains of subangular to angular quartz
(Fig. 5G). XRD data reveals that the grayer mudstone contains less
quartz and a higher percentage of clay minerals than the redder
mudstone.
Vertical, bifurcating traces (centimeter scale) are rare, whereas well-
developed slickensides are abundant (Fig. 5H), and autobrecciated fabric
occurs locally in which the clasts exhibit an angular shape. Planar laminae
of gypsum with vertically oriented crystals, 0.1–2 cm thick, locally occur
in these facies with sharp contacts against the massive mudstone.
Generally, the base of this facies grades upward from underlying
variegated mudstone facies with a transition zone consisting of complex
mudcracks and brecciated fabric whereas it is either sharply overlain by
the organic or variegated mudstone or more gradationally by the siltstone
facies.
T
ABLE
2.— Strontium isotope data from gypsum beds in the Wellington Formation.
Sample Identifier Gypsum Growth Habit
87/86
Sr (raw)
87/86
Sr (corrected) 2-Sigma Error
Gy26.5 displacive 0.70918 0.709186 0.0004
Gy26.5r displacive 0.7092375 0.7092435 0.00006
197.1b bottom growth 0.709244 0.70925 0.0001
KCl60.1 displacive 0.709094 0.7091 0.00009
Gy92.4 laminated 0.708682 0.708688 0.00008
197.1a bottom growth 0.709705 0.709711 0.00007
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 835
JSR
Interpretation.—The fine-grained and well-sorted texture and massive
bedding is most consistent with eolian deposition (Pye 1987; Tsoar and
Pye 1987). Typical far-traveled eolian ‘‘clayey’’ loess is of mud size and
enriched in mica and clay minerals (Junge 1969; Clemmensen 1979; Pye
1987). Abundant slickensides, ped-like features, and root traces record
pronounced pedogenic overprinting in a subaerial setting. The local
brecciated fabric and its angular texture suggests an in situ autoclastic
origin formed from wetting and drying (Benison and Goldstein 2001)
rather than erosion and traction transport of intraclasts, and no other
evidence of traction currents are present. The vertically aligned crystals in
the gypsum interbeds, however, suggest bottom precipitation from
occasional subaqueous conditions (Vandervoort 1997). This is evidence
for ephemeral saline surface waters.
Paleosols
Description.—Prominent, randomly oriented slickensides, blocky peds,
downwardly bifurcating vertical traces, and homogenized bedding occur
in discrete horizons throughout the core, and although most common to
the massive mudstone facies, are not limited to a single facies, and thus,
are summarized here. The slickensides are very pronounced, forming
shiny, subvertical planes in the mudstone that crosscut the entire core
(Fig. 5H) and are visible in outcrop forming high angles to bedding.
Blocky ped structures are also pervasive in these intervals. Microfabrics
include highly birefringent clay, exhibiting two preferred directions.
Intervals throughout the core exhibiting these features have a sharp upper
boundary and a gradational basal boundary. Rare downwardly
bifurcating traces occur, but gray reduction spots are more common.
XRD analyses indicate illite and smectite as the most prominent clay
minerals.
Interpretation.—The presence of (downwardly bifurcating) root traces,
abundant slickensides, and blocky peds record pedogenesis, specifically
development of Vertisols (Retallack 1990). Slickensides form from
shrinking and swelling of clay, particularly smectite during wetting and
drying (Retallack 1990), and the basal gradational boundary and sharp
upper boundary are consistent with pedogenesis. The structureless
bedding and lack of horizonation reflect pedoturbation as a result of
the shrinking and swelling (cf. Goebel et al. 1989). The oriented
microfabric is defined as clinobimasepic fabric (Retallack 1990) and
commonly occurs in Vertisols due to stress produced from shrinking and
swelling of clays (Nettleton and Sleeman 1985). Pedogenic overprinting is
most common in the massive mudstone facies and, subordinately,
variegated mudstone facies.
Geochemical Analysis
Whole-Rock and Trace-Element Geochemistry.—Whole-rock geochem-
istry of the Wellington provides information on potential provenance of
the sediment and chemical processes during deposition and early
diagenesis. In the following we compare the chemical data to several
reference compositions to ultimately test hypotheses of provenance but
also depositional processes. These reference datasets include: upper
continental crust (UCC; Taylor and McLennan 2001), average shale (SH;
Taylor and McClennan 1985), Quaternary loess (QL; Taylor and
McClennan 1985), upper Paleozoic loessite (UPL; Soreghan and
Soreghan 2007), Colorado Plateau crustal model (CPC; Condie and
Selverstone 1999); Ouachita flysch (OAU; Gleason et al. 1995; Sutton and
Land 1996; Totten et al. 2000), and roughly coeval Permian redbeds from
western Kansas (KA; Cullers 1994). The locations of these regional
datasets are shown in Figure 2.
The SiO
2
values (Fig. 7A) for the siltstone facies are higher (58 to 88.5
wt.%) and more variable than the mudstone facies (,58 to 60 wt.%); the
siltstone SiO
2
values are also higher than upper continental crust (UCC),
but similar to upper Paleozoic (UPL) and Quaternary loess (QL) values.
The mudstone facies samples are enriched in Al
2
O
3
compared to the
siltstone facies, UCC, UPL, and QL, but are similar to average shale (SH)
and Ouachita flysch (OAU). The difference in the ratio of SiO
2
/Al
2
O
3
between the mudstone and siltstone likely reflects grain-size effects rather
than differences in weathering or provenance owing to the higher
concentration of clay minerals in the mudstone (Taylor and McClennan
1985; Cullers 1994).
The crossplot of total alkali metal oxides (Na
2
O+K
2
O) vs. SiO
2
(Fig. 7B) shows a depletion relative to UCC in alkalis and no trend in the
mudstone facies, further reflective of a uniform composition of the
various mudstone facies. The siltstone facies displays a negative trend
(Fig. 7B), with values similar to UPL and QL. The Wellington mudstone
samples are relatively uniform and depleted in Na
2
O compared to the
siltstone samples when normalized to Al
2
O
3
,whereasthesiltstonefaciesis
more variable (Fig. 7C). The mudstone samples are again more similar to
SH and OUA, while the siltstone samples are more similar to QL, KA,
and UPL in the normalized plot. Na
2
O and K
2
O ratios in sediments are
typically controlled by source rock (Garrels and MacKenzie 1971), but
also can reflect weathering either prior to deposition or in situ (McLennan
et al. 1993; Qiao et al. 2009).
Differences in trace-element composition exist among the Wellington
facies. A plot of Th/Sc vs. Cr/Th (Fig. 7D) illustrates that the mudstone
samples uniformly lie between ratios for UCC (Th/Sc 51), which is
enriched in incompatible elements, and more mafic sources (Th/Sc 50.6;
Totten et al. 2000). In contrast, the siltstone samples exhibit a more
variable Th/Sc ratio, typical of a continental signature (Fig. 7D). The
siltstone samples plot similar to KA and UCC, whereas the mudstone
samples plot similar to the SH and OUA values.
Finally, there are a number of other differences in trace-element
composition of the mudstone compared to the reference data (Fig. 7E).
For example, the mudstone facies are enriched in Cr, Ni, and V compared
to UCC but depleted in Sr compared to UCC. The mudstone facies also
exhibits differences in trace-element pattern compared to average shale
(SH; Fig. 7E).
Detrital-Zircon Geochronology.—The Wellington Formation contains
zircons with U-Pb ages that range from Archean in age to ,350 Ma
(Appendix 1; Fig. 8). A large number of zircon grains (n514) fall into a
group that spans early Paleozoic (530–370) ages (Fig. 8). There are a few
grains in the age range of 570–740 Ma; however, the dominant mode are
grains (n531) of mostly Mesoproterozoic age that cluster within the
ranges of 900–1360 Ma. Finally, there is also a population of grains
(n511) that yield ages between 1610 and 1800 Ma and a few scattered
Archean ages (.2500 Ma, n54).
Provenance Interpretations of Geochemistry
The composition of the Wellington sediments differs significantly from
UCC. Multiple cross plots of the geochemistry data of the Wellington
sediments, including Na
2
O/Al
2
O
3
vs. K
2
O/Al
2
O
3
, Th vs. Sc, Th/Sc vs. Cr/
Th, and Th/U vs. Th, all show that the composition reflects mixed felsic
and mafic provenance, suggesting derivation from multiple sediment
sources.
The mudstone composition plotted on Na
2
O/Al
2
O
3
vs. K
2
O/Al
2
O
3
(Fig. 7C) and Th/Sc vs. Cr/Th (Fig. 7D) empirically suggests the
Ouachita Mountains (OAU) as a possible sediment source. The
composition of the Ancestral Rocky Mountains (ARM) is represented
in this study through derivative sediments, i.e., upper Paleozoic loess
(UPL; proximal) and Kansas shale (KA; distal) and is also similar to
the composition of the Wellington sediments. Overall, both KA, which
represents distal ARM sediment (Cullers 1994), and the Ouachita
836 J.M. GILES ET AL.
JSR
F
IG
.7.GeochemicalplotsofWellingtonmudstone(circles)andsiltstone(triangles)comparedtovariousgeochemicalreferencedatasets:UCCuppercontinental
crust; OUA Ouachita-derived sediments; KA Permian shales in Kansas; UPL upper Paleozoic loessite; SH average shale. See text for references. A) Crossplot of Al
2
O
3
vs.
SiO2, B) crossplot of Na
2
O+K
2
O vs. SiO
2
.C) Plot of Na
2
O/Al
2
O
3
vs. K
2
O/Al
2
O
3
. Mudstones plot between mafic and felsic compositions. Igneous, sedimentary, and
shale lines are from Garrels and MacKenzie (1971). D) Plot of Th/Sc vs. Cr/Th, with the bold line representing an idealized signature of a mixing line between granite and
MORB (mid-oceanic-ridge basalt) sources. The Wellington mudstone facies plots along the mixing line, while the Wellington siltstone facies trend toward a more felsic
composition. After Totten et al. (2000). E) Spider plot of average trace-element value of mudstone facies compared to various reference datasets.
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 837
JSR
Mountains detritus (OAU) both seem to best fit the relative proportion
of the trace elements in the Wellington mudstones (Fig 7E), particularly
when compared to average shale or upper continental crust. Although
the ARM contains higher silica content than the Ouachita Mountains
flysch, both are composed mainly of felsic-derived rock types (Cullers
1994; Totten et al. 2000) and exhibit similar compositions.
In contrast, the Wellington siltstone facies differs in composition from
the mudstone facies and trends toward a more felsic source. The
Wellington siltstones consistently plot similar to Paleozoic sediments
derived from the ARM, although a few siltstone samples overlap the
mudstone samples. Thus, overall the siltstone geochemistry signals
primary derivation from the ARM or its derivative sediments, with
either some secondary mixing with mudstone during pedogenic over-
printing or periodic sediment input from the Ouachita Mountains.
Another characteristic of the geochemical data is the uniform
composition of the different mudstone facies, which may indicate
sediment derivation from similar, well-mixed, distal source(s). Totten et
al. (2000) interpreted multiple sources for the Pennsylvanian Stanley
Group based on two distinct clusters in plots of Th vs. Sc and Th/Sc vs.
Cr/Th (Fig. 4 and 5 in Totten et al. 2000). In contrast, the Wellington
mudstone samples cluster tightly, typically between compositions of
felsic and mafic source rocks. Similarly, Cullers (1994) inferred a
homogeneous, well-mixed sediment source for distal mudstone deposits
compared to a heterogeneous, less-well-mixed source in more proximal
coeval deposits.
The detrital-zircon ages of the single sample analyzed are consistent
with interpretations of a mixed source. Although forming the most
abundant population, zircons with ages ranging from ,900 to 1300 Ma
(Fig. 8) are a common signature in upper Paleozoic deposits (Dickinson
and Gehrels 2003) that reflect contribution from Grenville basement and
do not necessarily suggest a unique source region for detrital grains in
the Wellington Formation. Grenville basement and derivative sediment
with grains of these ages lies both east (Becker et al. 2005) and south
(Gleason et al. 2007) of the study site. The population of grains in the
1600–1800 Ma range (Fig. 8) likely were derived from the Ancestral
Rocky Mountains or derivative deposits to the west of the study site,
where basement terranes (Yavapai–Mazatzal) of that age were exposed
throughout the Pennsylvanian and earliest Permian (Van Schmus et al.
1993), and zircons of these ages are a common component of slightly
older loessite deposits (M. Soreghan et al. 2002; Soreghan et al. 2008a).
Zircons of mid-Paleozoic age have been reported from terranes in the
Appalachian–Ouachita systems, and particularly terranes of Mexico and
Central America then situated south of the Ouachita system (Weber
et al. 2006; Martens et al. 2010; Soreghan and Soreghan 2013). This
suggests a south-southeast source direction for the mid-Paleozoic
detrital zircons.
DISCUSSION
Depositional Setting for the Wellington Formation
Since the original studies of the Wellington Formation, the deposi-
tional setting of this unit has been debated (Dunbar 1924; Raasch 1946;
Tasch 1964; Schultz 1975; Chaplin 2004; Hall 2004). Virtually all workers
have noted that there is definite evidence for continental deposition, but a
number of workers still favor a marine or marginal marine origin for
most of the Wellington, and attempt to reconcile the mixture of signals by
calling on either a tidal or sabkha environment or invoking numerous
transgressions and regressions at various temporal scales. Our facies,
petrographic, and geochemical data strongly suggest that the Permian
Wellington Formation of north-central Oklahoma formed in continental
environments, including saline and freshwater perennial lakes, playas,
wind-blown silt fields (loess), and soils. There is no diagnostic evidence
for marine or marginally marine paleoenvironments in the Anhydrite
sequence, the Otoe Member, and the Midco Member present in the KC-1
core and nearby outcrop.
The KC-1 core contains numerous paleosols, most of which are
inferred to be Vertisols. Paleosols in the Wellington were also
documented by Raasch (1946) and Hall (2004). Vertisols unequivocally
reflect extensive subaerial exposure and pedogenic overprinting of the
underlying sediment. If this underlying sediment is interpreted as marine,
or even intertidal, however, then superposition of these pedogenic
features necessitates numerous and very abrupt fluctuations in sea level.
For example, in Pennsylvanian cyclothems, paleosols commonly occur at
each cycle top, commonly on the 10 m (100–400 ky) scale in subsiding
cratonal regions (e.g., Heckel 1986; Davydov et al. 2010), whereas
paleosols occur on average every 3–5 m in the Wellington. In a lake basin,
however, abrupt changes of lake level, from relatively deep to dry, can
occur on the order of 10
4
years or less, even in very large tropical lakes
(e.g., Lake Victoria; Johnson et al. 1996).
Evaporites form in a wide range of environments, including marine,
marginal marine, and continental, which has obfuscated interpretations
of the Wellington Formation. The Sr isotope composition of the
gypsum, however, indicates a continental origin. The
87
Sr/
86
Sr compo-
sition of Permian seawater is well characterized and drops through the
early Permian (Korte et al. 2006) from ,0.7080 at the start of the
Permian to ,0.7070 in the Kungurian, then rises to 0.7077–0.7074
during the Artinskian. The data from the Wellington, averaging 0.7092,
indicate a (mainly) continental source. Given that some of the gypsum
records subaqueous formation, the most likely environment is a saline
perennial lake, whereas the gypsum that precipitated displacively likely
reflects a playa environment. We use the terms saline lake to refer to
continental basins with input of both groundwater and surficial water
from continental sources; playa to refer to continental basins with
F
IG
.8.CumulativeprobabilityplotofU-Pb
ages of detrital zircons from the outcrop sample
of the Midco Member of the Wellington
Formation. The curve represents the sum of
individual radiometric ages and associated errors
of the analyzed zircons. The significance of the
age peaks are discussed in the text.
838 J.M. GILES ET AL.
JSR
mainly meteoric groundwater input; and sabkha to refer to arid
shorelines with periodic input of surficial marine water (Yechieli and
Wood 2002).
Since the dolomite is intimately associated with the gypsum, then the
setting of the dolomite formation is more consistent with a continental
environment rather than a sabkha or lagoonal setting. Specifically, the
fine grain size and laminated character is consistent with a perennial,
saline lake in which the dolomite formed as a primary precipitate.
Dolomicrite, commonly interlaminated with siliciclastic material, is
common in the geologic record in units interpreted as offshore lake
facies (Desborough 1978; Deckker and Last 1988; Last 1990; Calvo et al.
1999). Indeed, the precipitation of gypsum could be one factor in
elevating the Mg ratio of the solution from which the dolomite ultimately
formed. A perennial-lake origin for the laminated dolomite is also
consistent with its co-occurrence with the organic-rich mudstone facies,
which further suggests low-oxygen, perennial lake conditions.
In addition to those attributes suggestive of a perennial lake, several
intervals in the study section display features more consistent with
periodic exposure in a playa environment, such as common desiccation
features, displacive gypsum, and mudstone displaying disrupted bedding
(autobrecciation and inferred efflorescent crusts; Spencer and Lowenstein
1990; Benison and Goldstein 2001). Although a number of these
sedimentary features are characteristic of a range of environments, their
interbedding with the inferred perennial-lake deposits, and the presence of
interbedded gypsum, with its continental chemical signature, suggests a
playa setting rather than sabkha, supratidal, or lagoonal setting.
The paleontologic evidence from the Wellington suggests a continental
signature for the study area with more brackish conditions to the north in
Kansas. The only fossils observed in the core are conchostracans found
on bedding planes of the laminated mudstone facies, which appear
pristine and in situ.BythePermian,conchostracanswerecommonin
continental environments (Tasch 1964; Swanson and Carlson 2002; Park
and Gierlowski-Kordesch 2007; Scott 2010); however, some conchos-
tracan species, commonly Triassic in age, may indicate brackish
conditions (Kozur and Weems 2010). In outcrops of the Wellington,
the ranges of fossils vary depending on location. Fossils reported in
Oklahoma (Table 3) outcrops indicate more freshwater conditions,
including conchostracans (Tasch 1961; Hall 2004), lungfish burrows
(Carlson 1999; Chaplin 2004), vertebrate ichnofauna (Swanson and
Carlson 2002), occasional stromatolites (Hall 2004), and insect beds
(Beckemeyer and Hall 2007). In contrast, fossils reported in Kansas
outcrops (Table 3) may indicate more brackish to marine conditions,
including eurypterids (Eurypterus sellardsi; Dunbar 1924), horseshoe
crabs (see below; Dunbar 1924; Hall 2004), stromatolites (Carlson 1968;
Hall 2004; Bolhar and Van Kranendonk 2007), Lingula (Dunbar 1924;
Hall 2004), and pelecypods (Dunbar 1924; Hall 2004). Stratigraphically,
the Oklahoma outcrops are reported as younger, including the Midco
through basal Billings Pool members, whereas the slightly older Kansas
outcrops include the upper Annelly Gypsum through the Carlton
Limestone (Midco equivalent) members (Table 3). Therefore, the
paleontology alone suggests that the stratigraphically older Wellington
fossils of Kansas indicate more brackish conditions, while the younger
Wellington fossils of Oklahoma suggest more continental conditions.
Finally, albeit negative evidence, the Wellington strata examined in this
study lack characteristics that generally define a sabkha or tidal
environment. For example, characteristic features such as, e.g., ‘‘chick-
en-wire’’ and enterolithic evaporites (gypsum or anhydrite), widespread
fenestral fabrics, flat-pebble intraclast conglomerates, and tepee struc-
tures (Warren and Kendall 1985) are rare to absent in the study strata. In
addition, sabkha deposits are typically interbedded with shallow subtidal
deposits (Schreiber 1978), whereas the Wellington lacks associated
normal marine deposits. Finally, no evidence occurs for the peritidal
‘‘trinity,’’ i.e., subtidal, intertidal, and supratidal subenvironments
(Warren and Kendall 1985), marked by facies and features such as
peloidal, oolitic, and skeletal grainstone, burrowed carbonate mudstone,
flaser bedding, and ripple cross-stratification.
Transport Processes of Wellington Formation Siliciclastics
The commonly accepted mode of transport for Pennsylvanian–Permian
clastics occurring in Oklahoma is fluvio-deltaic, with ultimate preserva-
tion in environments interpreted to range from fluvial and lacustrine to
deltaic and tidal flat (Twenhofel 1926; Wilson 1927; Fay 1964; Johnson
1990; Mazzullo 1999). These interpretations are based in part on evidence
such as the presence of cross-bedded sandstone, which occurs in the
Garber Sandstone immediately overlying the Wellington Formation
(Raasch 1946; Elrod 1980) and the presence of inferred marine evaporites
in the enclosing formations (Twenhofel 1926; Wilson 1927; Fay 1964;
T
ABLE
3.—General summary of paleontology found in the Wellington Formation. Note that brackish to marine environmental indicators are most common
in Kansas.
Fossil Type
Location (County)
Member
Interpreted Environment from
Source Source
Oklahoma Kansas
Conchostracans Noble and Kay Sumner, Sedgwick,
Harvey, Marion,
Dickinson
basal Billings Pool Continental Tasch 1964; Swanson and
Carlson 2002; Park and
Gierlowski-Kordesch
2007; Scott 2010; Kay
County Core
Lungfish burrows Noble none reported basal Billings Pool Continental Carlson 1999; Chaplin 2004
Insect beds Noble Dickinson basal Billings Pool (OK);
Carlton Limestone (KS)
Continental Beckemeyer and Hall 2007
Stromatolites Noble and Kay Dickinson basal Midco (OK); basal
Carlton Limestone (KS)
Brackish to Marine Hall 2004
Vertebrate ichonofauna Noble Dickinson basal Billings Pool Continental Swanson and Carlson 2002
Eurypterus sellardsi none reported Dickinson basal Carlton Limestone Brackish to Marine Dunbar 1924; Hall 2004
Horseshoe crabs
(Paleolimnus signatus)
none reported Dickinson basal Carlton Limestone Disagreement between
authors: Continental
or Brackish
Dunbar 1924; Babcock et al.
2000; Hall 2004
Pelecypods none reported Dickinson basal Carlton Limestone Brackish to Marine Dunbar 1924; Hall 2004
Lingula none reported Dickinson Annelly Gypsum/basal
Carlton Limestone
Brackish to Marine Dunbar 1924; Hall 2004
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 839
JSR
Johnson 1990). Some researchers, however, have postulated an eolian
origin for the mudstone in the Wellington and younger Flowerpot
formations (Raasch 1946; Yang 1985; respectively). In the Midco and
Otoe members of the study area, evidence for fluvio-deltaic deposition
does not occur: the siliciclastic strata exhibit a narrow (clay–silt) grain size
and lack evidence for erosional scouring or any channels. Cross-bedding
does occur in the Billings Pool Member in outcrop (above the study
interval) and oscillation ripple marks have been identified in thin
siltstones in the Midco Member (Hall 2004), yet the predominant
stratification type in the clastic facies is either laminated or massive.
The predominant facies in the Midco and Otoe members, the laminated
mudstone facies, reflects subaqueous deposition as indicated by very
continuous laminae and interlaminated dolomite and bottom-growth
gypsum. If the siliciclastic mud discharged as a strong overflow current
(hypopycnal flow) from one or more point sources in a density-stratified
lake the sediment would spread horizontally and ultimately sink as a
suspension deposit (e.g., Lake Turkana, Cohen 1989). This requires
considerable depth of the lake body to create the density gradient
required to buoy the sediment plume. The laminated mudstone facies,
however, contains numerous horizons of mudcracks, suggesting numer-
ous desiccation events. Thus, if the mud laminae were deposited as a
result of a series of overflow discharge events, then large-magnitude lake-
level fluctuations must have also been very abrupt and very common. If,
conversely, the mud discharged into the lake as a strong underflow
current (hyperpycnal flow) from one or more point sources, then distinct
sedimentary features should occur, e.g., laminae that exhibit grading, and
association with silty and at least fine sandy beds (Soyinka and Slatt
2008), which are absent.
A third possibility is that the clastic sediment was introduced into the
lake system via eolian processes. In the study section, grain-size averages
are in the very fine- to fine-silt range, typical for distal eolian dust (Pye
1987; Smalley et al. 2005), and dust deposition in lakes and playas is a
common phenomenon, particularly downwind of arid to semiarid regions
(e.g., southwestern U.S., Reheis et al. 1995; Lake Biwa, Xiao et al. 1999;
Lake Chad, Evans et al. 2004; man-made Lake Volta, Breuning-Madsen
and Awadzi 2005; northwest China, Qiang et al. 2007; Dead Sea, Haliva-
Cohen et al. 2012). In most of these modern systems, the eolian
component is of very fine silt to clay size and occurs as laminae in the lake
sediment or in more massive beds disturbed by biogenic reworking. Dust
transport and settling into a shallow, saline lake subject to occasional
desiccation, would produce mudcracked surfaces. Taking into consider-
ation the bulk of the evidence, including the lack of channels or sediment
thickening and coarsening indicative of point-source inputs of the
siliciclastic component to the lacustrine system, we favor eolian transport
as the chief mechanism of delivery of siliciclastic grains to the
depositional site. This process resulted in formation of the laminated
mudstone facies through subsequent suspension settling of the sediment
in a perennial lake that experienced periodic desiccation that formed the
millimeter scale laminae and mudcracks, respectively.
As noted above, the massive mudstone facies is also interpreted as
eolian in origin. The similar texture of the laminated and massive
mudstone, transitional contacts, and lack of erosional scours or lag
features between the facies suggest a similar transport mechanism. The
massive character and association with Vertisols could be interpreted to
reflect unconfined flow in which mud aggregates were carried in traction
(Wright and Marriott 2007). The massive mudstones in this case,
however, lack almost all of the common macroscopic attributes of such
traction-flow deposits as defined by Wright and Marriott, such as 1)
erosional bases, 2) intraclast lenses or lags, 3) related facies associations
(e.g., conglomeratic lenses, pedogenic carbonates), and 4) lenticular
stratification of interbedded calcrete clasts. Thus, we favor a mainly
eolian transport interpretation, in which the mud settled onto a playa or
shallow lake system, and was subsequently further homogenized
pedogenically. This still allows the possibility of occasional unconfined
flow events associated with storms to remobilize and locally transport
previously deposited mud (Alonso-Zarza et al. 2009). This also does not
eliminate the presence of aggregates, and the microscopic observation of
quartz clasts floating in mud matrix (Fig. 5G) suggests that aggregates
might have existed. However, we suggest that if the mud did form
aggregates, these were transported via eolian processes and might be
similar to Quaternary eolian deposits documented in Australia (parna).
Butler (1956) suggested that these massive, clayey, pedogenically altered
deposits (1–3 m thick) were formed by wind transport of silt-size clasts of
clay aggregates, although this process remains obscure (Hesse and
McTainsh 2003).
Finally, the siltstone facies are interpreted as loess(ite) deposits that
accumulated over the playa system during phases when no standing water
occurred. The massive texture, well-sorted silt grain size, and provenance
are all consistent with an eolian origin, and no features occur consistent
with traction flow, whether confined or unconfined (e.g., lateral-accretion
surfaces, or planar or ripple cross lamination). The provenance, reflected
in both the geochemical and detrital zircon attributes, indicates a mixed
source that reflects regions both east and west of the study area,
consistent with a seasonally changing wind regime.
Wellington Lake Model and Regional Paleoclimatic Implications
The climate-sensitive sediments composing arid, closed-basin lakes
record high-frequency variations in regional short-term climates and are
useful for paleoclimate interpretations (Smith et al. 1983; Olsen 1986;
Smith and Bischoff 1997; Lowenstein et al. 1999; Lowenstein et al.
2003). No evidence exists for major changes in tectonic regime during
deposition of the Wellington Formation; thus, deposition was chiefly
influenced by regional climate changes and should be considered an
‘‘optimal’’ reflection of paleoclimate (Smith and Bischoff 1997; Low-
enstein et al. 2003). The facies in the Wellington Formation appear to
reflect at least two different temporal scales of climate change. First, the
facies stacking suggest repeated, upwardly shallowing events (detailed
below). Second, a more detailed examination of the facies transitions
suggests that, in the upwardly shallowing successions, small-scale
fluctuations indicative of higher-frequency wetting and drying occurred,
evinced by juxtaposition and overprinting of facies and features,
including: 1), the syncopated pattern of facies, for example the organic
and laminated mudstone facies; 2) specific features, such as Vertisols
containing slickensides; 3) alternating laminae of evaporites (dry) and
clastics (wet); and 4) complex mudcracks reflecting seasonality. In
addition, the presence of conchostracans indicates seasonality, as their
eggs are laid in ephemeral water bodies (Tasch 1958; Swanson and
Carlson 2002); similarly, lungfish, reported from the Wellington,
aestivate in burrows during dry episodes (Carlson 1968; Hasiotis et al.
1993; Swanson and Carlson 2002).
Vertical facies transitions in the study section indicate a preferred
vertical stacking pattern of, from base to top: (1) organic mudstone with
dolomite, (2) laminated mudstone with gypsum, (3) dark red or gray
massive mudstone, and (4) fine-grained siltstone. Using the organic
mudstone as a base for each, the study interval records at least 30
deepening and subsequent shallowing episodes (Fig. 3) that likely reflect
longer-term climatic change during Wellington Formation deposition.
Perennial Freshwater Lake Stage.—This stage represents the ‘‘flooding
stage’’ and the most humid interval as groundwater, direct precipitation,
and related runoff filled the basin and deposited the organic mudstone
facies (Fig. 9). It is typical in tropical lake systems that direct
precipitation and evaporation dominate the local hydrology, for example,
even in modern Lake Tanganyika, the second largest freshwater lake in
the world by volume, the main contribution to water input is direct
840 J.M. GILES ET AL.
JSR
F
IG
. 9.—Depositional model for the Welling-
ton Formation. Each stage depicted in the
diagram is generalized; high-frequency fluctua-
tions likely occurred, creating heterogeneous
facies. Perennial-lake stage is represented by the
organic mudstone and laminated dolomite. (2)
Perennial lake, but with increasing evaporation,
causing an overall decrease in water level and
formation of the variegated mudstone facies and
thicker gypsum and dolomite units, though
seasonality and short-term fluctuations allowed
occasional return of deposition of organic
mudstone. (3) Playa or mudflat stage, during
which desiccation features are common and the
variegated (or rarely the organic) mudstone was
pedogenically overprinted, creating Vertisols.
Additional eolian transport of clayey loess
coupled with pedogenesis formed the massive
mudstones. (4) Loess deposition.
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 841
JSR
precipitation (Nicholson 1999). This stage represents in aggregate ,25%
(by thickness) of the study interval, and thus, a significant part of the
paleogeography. The lake was likely stratified and possibly anoxic,
enhancing sulfate reduction and dolomite formation, and preservation of
laminations and organics (Trudinger et al. 1985; Last 1990; Machel 2004;
Sanz-Montero et al. 2009). Seasonality or short-term changes in
hydrology likely caused fluctuations in salinity and water level as
evidenced by variable thicknesses of dolomite deposits, some with
desiccation features, and rare gypsum laminations in the organic
mudstone facies. Rare periods of increased salinity and evaporation
would trigger subaqueous gypsum growth on the lake bottom and within
the sediment, analogous to other examples of evaporites forming
contemporaneously with high rates of organic deposition along a lateral
salinity gradient (e.g., Nury and Schreiber 1997). Arid episodes may have
also resulted in eolian transport of silt-size clastics into the lake, creating
rare organic-poor (lighter-colored) silty laminae.
Perennial Saline Lake Stage.—This stage represents increasing aridity
leading to greater evaporation rates and higher salinity, and consequent
deposition of the laminated variegated mudstone (Fig. 9). Although this
stage represents a shallower, less anoxic lake, evidence for strong
seasonality and a fluctuating water chemistry includes: 1) local dolomite
laminations, 2) abundant desiccation features, including complex
mudcracks penetrating throughout each mudstone bed and at the top
of the interbedded dolomite beds, 3) small pedogenic slickensides, 4)
brecciated layers and chaotic bedding thinly interbedded between
laminated intervals, and 5) variegated colors (Chaplin 2004). Pulses of
eolian silt and clay entered the lake and formed the discrete silty
laminations. Most of the gypsum in the core formed during this stage,
with both subaqueous vertical crystal precipitates and intra-sediment
precipitates. The presence of both of these forms of gypsum vertically
juxtaposed further indicates abrupt fluctuations in water level (Benison
and Goldstein 2001). Conchostracans occur during this stage and were
able to tolerate the high-stress environment of a saline lake (Tasch 1961a),
but they occur in beds lacking evaporites and thus probably represent
periods of slightly decreased salinity.
Ephemeral (Playa) Stage.—During this stage, much of the water
covering the basin evaporated, exposing the laminated variegated
mudstone (Fig. 9). Transitional contacts between the variegated mudstone
and overlying massive mudstone (Vertisols) facies support a similar (initial
eolian) origin for these clastics, with the massive mudstone reflecting both
ultimate deposition as loess, rather than subaqueous settling, and
pedogenic overprinting. Several authors suggest that the exposed surfaces
in saline lake environments experience almost constant pedogenic
overprinting (Bowler 1986; Smoot and Lowenstein 1991; Gierlowski-
Kordesch and Rust 1994). Seasonal wetting and drying of mudstones
during pedogenesis led to development of Vertisols. The geochemical
similarity between the massive and laminated mudstones lends additional
evidence for the genetic relationship between these two facies.
Loessitic Stage.—The final stage is represented by eolian deposition of
the siltstone facies, reflecting the most prolonged arid phases (Fig. 9).
Rare pedogenic overprinting in the siltstone (indicated by root traces)
may reflect periods of decreased sediment influx. Trapping of loess
requires moist ground and thus, seasonal wetness (Pye 1987). As saline
groundwater evaporated from these silt deposits, efflorescent salts
formed, resulting in the patchy fabric, and chaotic bedding, analogous
to modern examples in Saline Valley and Death Valley (Gierlowski-
Kordesch and Rust 1994; Smoot and Castens-Seidell 1994).
Throughout deposition of the Wellington Formation, an even longer-
term shift in climate is inferred based on a significant change in the relative
proportion of facies between the Midco Member and underlying Otoe
Member and Anhydrite sequence. The lower part of the interval (,50 m,
Otoe Member and Anydrite sequence; Fig. 3) contains a substantially
larger percentage of facies reflecting arid conditions, specifically the very
fine siltstone, red massive mudstone, laminated variegated mudstone, and
gypsum beds. In contrast, the upper part of the interval (,40 m, Midco
Member; Fig. 3) is composed of a markedly higher percentage of facies
reflecting more humid conditions, including the gray massive mudstone,
organic mudstone, and dolomite. Thus, the study section overall reflects a
shift from more arid conditions in the Otoe and Anhydrite members to
more humid (relative) conditions in the Midco Member (Fig. 3).
This model suggests a generally semiarid but highly seasonal climate
for the mid-continent in late Early Permian time. Many have inferred that
zonal circulation in western and central tropical Pangea during the
Pennsylvanian shifted to increasingly monsoonal and thus, increasingly
seasonal climate in the Permian (Parrish 1993; Gibbs et al. 2002; G.
Soreghan et al. 2002; Tabor and Montan˜ez 2002). The provenance data
from the Wellington suggest that sediment emanated from both the west-
northwest (ARM) and south-southeast (Ouachita Mountains; in paleo-
coordinates). This implies northwesterly and south-southeasterly winds
(Fig. 9) if: 1) the Wellington sediments were eolian-sourced and 2) the
dust source remained proximal to the original basement source. Relating
these interpretations to the depositional model of the Wellington, the
laminated and massive (paleosols) mudstones represent more humid
periods associated with wetter times accompanied by south (easterlies),
whereas loess deposits (very fine siltstone facies) represent drier
conditions with increased aridity and seasonal westerlies.
The temporal and lateral extent of this lake system is difficult to
constrain based on the core and associated outcrops of this study.
Delineating the size of this inferred lake system is also hindered by the
lack of other detailed sedimentologic studies of the Wellington that would
define potential lacustrine deposits. Nevertheless, previous workers have
correlated individual marker beds in the Midco and Otoe members of the
Wellington between northern Oklahoma and central Kansas, over 220 km
(Fig. 1; Raasch 1946; Hall 2004). Employing the simplifying assumption
that this outcrop belt represents the minimal diameter of a circular lake,
this suggests a single lake of over 37,000 km
2
. This speculative but
supportable calculation paints a paleogeographic portrait of the Early
Permian Midcontinent that differs substantially from previous interpre-
tations and provides testable hypotheses for future studies.
CONCLUSIONS
1. The lithologic and geochemical data presented in this study, coupled
with data from earlier studies, suggest that the Lower Permian
Wellington Formation of Oklahoma records continental (lacustrine
and loess) deposition rather than marginal marine (sabkha, tidal).
The chemical facies forms only 10% of the Wellington and comprises
1) predominantly fine-grained laminated dolomite interlaminated
with organic-rich mudstone that records subaqueous suspension
deposition and 2) gypsum that consists of several primary morphol-
ogies reflecting deposition in subaqueous and shallow subsurface
(displacive) environments, but exhibit Sr isotope compositions
inconsistent with Permian seawater. Siliciclastic facies make up most
of the Wellington and reflect both subaqueous suspension and eolian
deposition but overall eolian transport. The clastic facies shows no
indication of tidal or sabkha features. Common overprinting by
Vertisol pedogenic features also points to a continental setting.
2. Deposition of the Wellington Formation occurred in a range of
related environments that included perennial freshwater lacustrine,
perennial saline lacustrine, ephemeral lacustrine (playa), and loess
accumulations. These fluctuations suggest strong climatic control
on the facies successions on various scales. Seasonality in climate is
suggested by alternations of laminated sediment, Vertisols, and
842 J.M. GILES ET AL.
JSR
fossils (e.g., conchostraca). Longer-term climatic change is indicat-
ed by facies successions that reflect transitions from a perennial
freshwater lake with primary dolomite growth to a shallower, saline
lake with subaqueous gypsum growth and laminated mudstone to a
dry saline pan with development of mudcracks (desiccation stage),
pedogenic overprinting, and loess accumulation.
3. Geochemical and detrital zircon data indicate a mixed provenance
for the volumetrically predominant siliciclastic component, with
both south-southeastern (Ouachita and possibly Mexico) and
western (ARM) sources.
4. Evidence from this study is consistent with the interpretations of
Benison and Goldstein (2001) suggesting that a large area of
shallow lakes existed in the Midcontinent during the Permian.
Other coeval deposits should be re-examined for possible alternate
interpretations, which could fundamentally shift our view of the
Permian in the Midcontinent.
ACKNOWLEDGMENTS
We thank the Oklahoma Geological Survey, especially J. Chaplin, for
logistical support in core analysis. G. Morgan, A. Madden, and R.
Turner assisted in chemical and mineralogical analysis. G. Augsburger
(OU) conducted the detrital zircon analysis, and G. Gehrels provided
support in the LaserChron Facility at the University of Arizona. T.
Rasbury, SUNY Stony Brook, provided Sr isotopic data. We also thank
T. Ruble at Hubble Geochemical Services, who provided complimentary
organic geochemical analyses. J. Pack thanks her family for their
dedication and support. We thank E. Gierlowski-Kordesch and an
anonymous reviewer for substantially improving an earlier version of this
manuscript, although they do not necessarily agree with all of our
interpretations. Funding for this project was provided by the U.S.
National Science Foundation (EAR-0746042 and EAR-1053018). Any
opinions, findings, and conclusions or recommendations expressed in this
material are those of the author(s) and do not necessarily reflect the view
of the National Science Foundation.
Appendix 1 contains the U-Pb geochronologic analyses of detrital zircons
in the Wellington Formation siltstone and is available from JSR’s Data
Archive: http://sepm.org/pages.aspx?pageid5229.
REFERENCES
A
LGEO
, T.J.,
AND
H
ECKEL
, P.H., 2008, The Late Pennsylvanian Midcontinent Sea of
North America: a review: Palaeogeography, Palaeoclimatology, Palaeoecology,
v. 268, p. 205–221.
A
LLEN
, R.A.,
AND
C
OLLINSON
, J.D., 1986, Lakes, in Reading, H.G., ed., Sedimentary
Environments and Facies: Oxford, U.K., Blackwell Scientific Publications, p. 63–94.
A
LONSO
-Z
ARZA
, A.M., Z
HAO
, Z., S
ONG
, C., L
I
, J., Z
HANG
, J., M
ARTI
´N
-P
E
´REZ
, A., M
ARTI
´N
-
G
ARCI
´A
, R., W
ANG
, X., Z
HANG
, Y.,
AND
Z
HANG
, M., 2009, Mudflat/distal fan and
shallow lake sedimentation (upper Vallesian–Turolian) in the Tianshui Basin, Central
China: evidence against the late Miocene eolian loess: Sedimentary Geology, v. 222,
p. 42–51.
B
ABCOCK
, L.E., D
ANIEL
, F.M.,
AND
R
ONALD
, R.W., 2000, Paleolimulus, an early limuline
(Xiphosurida), from Pennsylvanian through Permian Lagersttten of Kansas and
taphonomic comparison with modern Limulus: Lethaia, v. 33, p. 129–141.
B
ECKEMEYER
, R.J.,
AND
H
ALL
,J.D.,2007,TheentomofaunaoftheLowerPermianfossil
insect beds of Kansas and Oklahoma, USA: African Invertebrates, v. 48, p. 23–39.
B
ECKER
, T.P., T
HOMAS
, W.A., S
AMSON
, S.D.,
AND
G
EHRELS
, G.E., 2005, Detrital zircon
evidence of Laurentian crustal dominance in the lower Pennsylvanian deposits of the
Alleghanian clastic wedge in eastern North America: Sedimentary Geology, v. 182,
p. 59–86.
B
ENISON
, K.C.,
AND
G
OLDSTEIN
, R.H., 2000, Sedimentology of ancient saline pans: an
example from the Permian Opeche Shale, Williston Basin, North Dakota, U.S.A.:
Journal of Sedimentary Research, v. 70, p. 159–169.
B
ENISON
, K.C.,
AND
G
OLDSTEIN
, R.H., 2001, Evaporites and siliciclastics of the Permian
Nippewalla Group of Kansas, USA: a case for non-marine deposition in saline lakes
and saline pans: Sedimentology, v. 48, p. 165–188.
B
ENISON
, K.C.
AND
G
OLDSTEIN
, R.H., 2002, Recognizing acid lakes and groundwaters in
the rock record: Sedimentary Geology, v. 151, p. 177–185.
B
ENISON
, K.C., B
OWEN
, B.B., O
BOH
-I
KUENOBE
, F.E., J
AGNIECKI
,E.A.,L
ACLAIR
, D.A.,
S
TORY
, S.L., M
ORMILE
, M.R.,
AND
H
ONG
,B.Y.,2007,Sedimentologyofacidsaline
lakes in southern western Australia: newly described processes and products of an
extreme environment: Journal of Sedimentary Research, v. 77, p. 366–388.
B
ERG
,J.A.,1977,PetrographyandBromineGeochemistryoftheHutchinsonSalt
Member of the Wellington Formation in Ellsworth County, Kansas: a paleoenviron-
mental analysis [unpublished M.S. thesis]: University of Kansas, Lawrence, 74 p.
B
LAKEY
,R.C.,2007,CarboniferousPermianpaleogeographyoftheassemblyof
Pangaea, in Wong, T.E., ed., Proceedings of the XVth International Congress on
Carboniferous and Permian stratigraphy, Utrecht, 10–16 August 2003: Amsterdam,
Royal Dutch Academy of Arts and Sciences, p. 443–456.
B
OARDMAN
, D.R.I., 1999, Changes in patterns of cyclicity in Upper Carboniferous
through Lower Permian (Virgilian–Sakmarian) depositional sequences in the North
American Midcontinent [abstract], in Merriam, D.F., ed., Transactions of the 1999
American Association of Petroleum Geologists Midcontinent Section Meeting:
Lawrence, Kansas, Kansas Geological Survey Open-File Report 99-28, p. 86.
B
OLHAR
, R.,
AND
V
AN
K
RANENDONK
, M., 2007, A non-marine depositional setting for
the northern Fortescue Group, Pilbara Craton, inferred from trace element
geochemistry of stromatolitic carbonates: Precambrian Research, v. 155, p. 229–250.
B
OWLER
, J.M., 1986, Spatial variability and hydrologic evolution of Australian lake
basins: analogue for Pleistocene hydrologic change and evaporite formation:
Palaeogeography, Palaeoclimatology, Palaeoecology, v. 54, p. 21–41.
B
REUNING
-M
ADSEN
, H.,
AND
A
WADZI
, T.W., 2005, Harmattan dust deposition and
particle size in Ghana: Catena, v. 63, p. 23–38.
B
UTLER
,B.E.,1956,Parna:anaeolianclay:AustralianJournalofScience,v.18,p.145
151.
C
ALVO
, J.P., B
LANC
-V
ALLERON
, M.M., R
ODRIGUEZ
-A
RANDA
, J.P., R
OUCHY
, J.M.,
AND
S
ANZ
, M.E., 1999, Authigenic clay minerals in continental evaporitic environments, in
International Association of Sedimentologists, Special Publication 27, p. 129–151.
C
ARLSON
,K.J.,1968,TheskullmorphologyandestivationofburrowsofthePermian
lungfish, Gnathorhiza serrata: Journal of Geology, v. 76, p. 641–663.
C
ARLSON
, K.J., 1999, Crossotelos, an Early Permian nectridean amphibian: Journal of
Vertebrate Paleontology, v. 19, p. 623–631.
C
ARR
, T.R., 1982, Log-linear models, Markov chains and cyclic sedimentation: Journal
of Sedimentary Petrology, v. 52, p. 905–912.
C
ARTER
, L.S., K
ELLEY
, S.A., B
LACKWELL
, D.D.,
AND
N
AESER
, N.D., 1998, Heat flow and
thermal history of the Anadarko Basin, Oklahoma: American Association of
Petroleum Geologists, Bulletin, v. 82, p. 291–316.
C
HAPLIN
, J.R., 1988, Lithostratigraphy of Lower Permian rocks in Kay County, north-
central Oklahoma, and their stratigraphic relationships to lithic correlatives in Kansas
and Nebraska, in Morgan, W.A., and Babcock, J.A., eds., Permian Rocks of the
Midcontinent: Tulsa, Mid-Continent Section, SEPM, Special Publication 1, p. 79–
111.
C
HAPLIN
, J.R., 2004, Core drilling and stratigraphic analysis of Lower Permian rocks,
northern Oklahoma shelf, Kay County, Oklahoma: Oklahoma Geological Survey,
Special Publication 2004-1, 173 p.
C
HILINGAR
, G.V., 1957, Classification of limestones and dolomites on basis of Ca/Mg
ratio: Journal of Sedimentary Petrology, v. 27, p. 187–189.
C
LEMMENSEN
, L., 1979, Triassic lacustrine red-beds and palaeoclimate: the ‘‘Buntsand-
stein’’ of Helgoland and the Malmros Klint Member of East Greenland: Geologische
Rundschau, v. 68, p. 748–774.
C
OHEN
, A., 1989, Facies relationships and sedimentation in large rift lakes and
implications for hydrocarbon exploration: examples from Lakes Turkana and
Tanganyika: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 70, p. 65–80.
C
ONDIE
, K.C.,
AND
S
ELVERSTONE
,J.,1999,ThecrustoftheColoradoPlateau:newviews
of an old arc: Journal of Geology, v. 107, p. 387–397.
C
ROWELL
, J.C., 1999, Pre-Mesozoic Ice Ages: Their Bearing on Understanding the
Climate System: Geological Society of America, Memoir 192, 112 p.
C
ULLERS
, R.L., 1994, The controls on the major and trace element variation of shales,
siltstones, and sandstones of Pennsylvanian–Permian age from uplifted continental
blocks in Colorado to platform sediment in Kansas, USA: Geochimica et
Cosmochimica Acta, v. 58, p. 4955–4972.
D
AVYDOV
, V.I., C
ROWLEY
, J.L., S
CHMITZ
, M.D.,
AND
P
OLETAEV
,V.I.,2010,High-
precision U-Pb zircon age calibration of the global Carboniferous time scale and
Milankovitch-band cyclicity in the Donets Basin, eastern Ukraine: Geochemistry,
Geophysics, Geosystems, v. 11, no. Q0AA04, 22 p.
D
ECKKER
, P.D.,
AND
L
AST
, W.M., 1988, Modern dolomite deposition in continental,
saline lakes, western Victoria, Australia: Geology, v. 16, p. 29–32.
D
ESBOROUGH
, G.A., 1978, A biogenic–chemical stratified lake model for the origin of oil
shale of the Green River Formation: an alternative to the playa-lake model:
Geological Society of America, Bulletin, v. 89, p. 961–971.
D
E
V
OTO
,R.H.,1980,PennsylvanianstratigraphyandhistoryofColorado,inKent,
H.C., and Porter, K.W., eds., Colorado Geology: Rocky Mountain Association of
Geologists, 1980 Symposium, p. 37–50.
D
ICKINSON
, W.R.,
AND
G
EHRELS
, G.E., 2003, U-Pb ages of detrital zircons from Permian
and Jurassic eolian sandstones of the Colorado Plateau, USA: Paleogeographic
implications: Sedimentary Geology, v. 163, p. 29–66.
D
ICKINSON
, W.R.,
AND
G
EHRELS
, G.E., 2008, Sediment delivery to the Cordilleran
foreland basin: insights from U-Pb ages of detrital zircons in Upper Jurassic and
Cretaceous strata of the Colorado Plateau: American Journal of Science, v. 308,
p. 1041.
D
RIESE
, S.,
AND
D
OTT
, R., 1984, Model for sandstone–carbonate ‘‘cyclothems’’ based on
upper member of Morgan Formation (Middle Pennsylvanian) of northern Utah and
Colorado: American Association of Petroleum Geologists, Bulletin, v. 68, p. 574–597.
D
UNBAR
,C.,1924,KansasPermianinsects,Part1,thegeologicoccurrenceandthe
environment of the insects: American Journal of Science, v. 7, p. 171–209.
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 843
JSR
D
UNBAR
, C., B
AKER
, A., C
OOPER
, G.A., K
ING
, P., M
C
K
EE
, E., M
ILLER
, A., M
OORE
, R.,
N
EWELL
, N., R
OMER
, A.,
AND
S
ELLARDS
, E., 1960, Correlation of the Permian
formations of North America: Geological Society of America, Bulletin, v. 71, p. 1763.
E
LROD
, D., 1980, A Geochemical and Petrographic Survey of the Wellington
Formation, North-Central Oklahoma [unpublished M.S. thesis]: Oklahoma State
University, Stillwater, 100 p.
E
VANS
, R., J
EFFERSON
, I., K
UMAR
, R., O’H
ARA
-D
HAND
, K.,
AND
S
MALLEY
, I., 2004, The
nature and early history of airborne dust from North Africa; in particular the Lake
Chad basin: Journal of African Earth Sciences, v. 39, p. 81–87.
F
AY
,R.,1964,TheBlaineandRelatedFormationsofNorthwesternOklahomaand
Southern Kansas: Oklahoma Geological Survey, Bulletin 98, 238 p.
F
IELDING
, C., F
RANK
, T.,
AND
I
SBELL
, J., 2008, The late Paleozoic ice age: a review of
current understanding and synthesis of global climate patterns, in Geological Society
of America, Special Paper 441, p. 343–354.
F
ISCHBEIN
, S., J
OECKEL
, R.,
AND
F
IELDING
,C.,2009,Fluvialestuarinereinterpretationof
large, isolated sandstone bodies in epicontinental cyclothems, Upper Pennsylvanian,
northern Midcontinent, USA, and their significance for understanding late Paleozoic
sea-level fluctuations: Sedimentary Geology, v. 216, p. 15–28.
F
OLK
,R.L.,1973,CarbonatepetrographyinthePost-SorbianAge:JohnsHopkins
University, Studies in Geology, v. 21, p. 118–158.
G
ARCI
´A
D
EL
C
URA
, M., C
ALVO
, J.P., O
RDONEZ
, S., J
ONES
, B.F.,
AND
C
AN
˜AVERAS
, J.C.,
2001, Petrographic and geochemical evidence for the formation of primary, bacterially
induced lacustrine dolomite: La Roda ‘‘white earth’’(Pliocene, central Spain):
Sedimentology, v. 48, p. 897–915.
G
ARRELS
, R.,
AND
M
ACKENZIE
, F., 1971, Evolution of Sedimentary Rocks: New York,
W.W. Norton and Company, 397 p.
G
EHRELS
, G.E., V
ALENCIA
, V.A.,
AND
R
UIZ
,J.,2008,Enhancedprecision,accuracy,
efficiency, and spatial resolution of U-Pb ages by laser ablation–multicollector–
inductively coupled plasma–mass spectrometry: Geochemistry Geophysics Geosys-
tems, v. 9, no. Q03017, p. 1–13.
G
IBBS
, M., R
EES
, P., K
UTZBACH
, J., Z
IEGLER
, A., B
EHLING
, P.,
AND
R
OWLEY
, D., 2002,
Simulations of Permian climate and comparisons with climate sensitive sediments:
Journal of Geology, v. 110, p. 33–55.
G
IBERT
, L., O
RTI
´
, F.,
AND
R
OSELL
, L., 2007, Plio-Pleistocene lacustrine evaporites of the
Baza Basin (Betic Chain, SE Spain): Sedimentary Geology, v. 200, p. 89–116.
G
IERLOWSKI
-K
ORDESCH
, E.,
AND
R
UST
,B.,1994,TheJurassicEastBerlinFormation,
Hartford Basin, Newark Supergroup (Connecticut and Massachusetts); a saline lake–
playa–alluvial plain system, in Renaut, R.W., and Last, W.M., eds., Sedimentology
and Geochemistry of Modern and Ancient Saline Lakes: SEPM, Special Publication
50, p. 249–265.
G
ILBERT
, M., 1992, Speculations on the origin of the Anadarko Basin, in Mason, R., ed.,
Basement Tectonics: International Basement Tectonics Association, Publication 7,
p. 195–208.
G
ILBERT
, M., 2002, Estimating topographic heights in the Permian Wichita Mountains,
Oklahoma [abstract]: Geological Society of America, Abstracts with Programs, v. 34,
p. 472.
G
LEASON
, J.D., P
ATCHETT
, P.J., D
ICKINSON
, W.R.,
AND
R
UIZ
,J.,1995,Ndisotopic
constraints on sediment sources of the Ouachita–Marathon fold belt: Geological
Society of America, Bulletin, v. 107, p. 1192–1210.
G
LEASON
, J.D., G
EHRELS
, G.E., D
ICKINSON
, W.R., P
ATCHETT
, P.J.,
AND
K
RING
, D.A.,
2007, Laurentian sources for detrital zircon grains in turbidite and deltaic sandstones
of the Pennsylvanian Haymond Formation, Marathon assemblage, west Texas, USA:
Journal of Sedimentary Research, v. 77, p. 888–900.
G
OEBEL
, K.A., B
ETTIS
, E.A., III,
AND
H
ECKEL
, P.H., 1989, Upper Pennsylvanian paleosol
in Stranger Shale and underlying Iatan Limestone, southwestern Iowa: Journal of
Sedimentary Petrology, v. 59, p. 224–232.
G
OLONKA
, J.A.,
AND
F
ORD
, D., 2000, Pangean (Late Carboniferous–Middle Jurassic)
paleoenvironment and facies: Palaeogeography, Palaeoclimatology, Palaeoecology,
v. 161, p. 1–34.
H
ALIVA
-C
OHEN
, A., S
TEIN
, M., G
OLDSTEIN
, S.L., S
ANDLER
, A.,
AND
S
TARINSKY
, A., 2012,
Sources and transport routes of fine detritus material to the Late Quaternary Dead
Sea basin: Quaternary Science Reviews, v. 50, p. 55–70.
H
ALL
, J., 2004, Depositional Facies and Diagenesis of the Carlton Member (Kansas)
and the Midco Member (Oklahoma) of the Wellington Formation (Permian, Sumner
Group) [unpublished M.S. thesis]: Wichita State University, Wichita, Kansas, 112 p.
H
ARDIE
, L.A., S
MOOT
, J.P.,
AND
E
UGSTER
, H.P., 1978, Saline lakes and their deposits; a
sedimentological approach, in Matter, A.T., ed., Modern and Ancient Lake
Sediments: Oxford, U.K., International Association of Sedimentologists, Special
Publication, no. 2, p. 7–41.
H
ASIOTIS
, S.T., M
ITCHELL
, C.E.,
AND
D
UBIEL
, R.F., 1993, Application of morphologic
burrow interpretations to discern continental burrow architects: lungfish or crayfish?
Ichnos, v. 2, p. 315–333.
H
ECKEL
,P.H.,1986,Sea-levelcurveforPennsylvanianeustaticmarinetransgressive
regressive depositional cycles along Midcontinent outcrop belt, North America:
Geology, v. 14, p. 330–334.
H
ECKEL
, P.H., 2007, Advances in interpretive sedimentology and stratigraphy associated
with study of Pennsylvanian glacial–eustatic cyclothems: Proceedings of the XVth
International Congress on Carboniferous and Permian Stratigraphy, p. 23–34.
H
ESSE
, P.P.,
AND
M
C
T
AINSH
,G.M.,2003,Australiandustdeposits:modernprocesses
and the Quaternary record: Quaternary Science Reviews, v. 22, p. 2007–2035.
J
OHNSON
, K., 1990, Hydrogeology and Karst of the Blaine gypsum–dolomite Aquifer,
Southwestern Oklahoma: Oklahoma Geological Survey, Special Publication 90-5,
31 p.
J
OHNSON
, K., A
MSDEN
, T., D
ENISON
, R., D
UTTON
, S., G
OLDSTEIN
, A., R
ASCOE
, B.,
S
UTHERLAND
, P.,
AND
T
HOMPSON
,D.,1989,GeologyofthesouthernMidcontinent:
Oklahoma Geological Survey, Special Publication 89-2, p. 12–20.
J
OHNSON
, K., N
ORTHCUTT
, R., H
INSHAW
, G.,
AND
H
INES
, K., 2001, Geology and
petroleum reservoirs in Pennsylvanian and Permian rocks of Oklahoma: Oklahoma
Geological Survey, Circular 104, 20 p.
J
OHNSON
, S., 1989, Significance of loessite in the Maroon Formation (Middle
Pennsylvanian to Lower Permian), Eagle Basin, northwest Colorado: Journal of
Sedimentary Petrology, v. 59, p. 782–791.
J
OHNSON
, T.C., S
CHOLZ
, C.A., T
ALBOT
, M.R., K
ELTS
, K., R
ICKETTS
, R., N
GOBI
, G.,
B
EUNING
, K., S
SEMMANDA
, I.,
AND
M
C
G
ILL
, J., 1996, Late Pleistocene desiccation of
Lake Victoria and rapid evolution of cichlid fishes: Science, v. 273, p. 1091–1093.
J
ONES
, C.L., 1965, Petrography of evaporites from the Wellington Formation near
Hutchinson, Kansas: United States Geological Survey, Bulletin B1201, p. 1–70.
J
UNGE
, C.E., 1969, Comments on the concentration of size distribution measurements of
atmospheric aerosols and a test of the theory of self-preserving size distribution:
Journal of Atmospheric Science, v. 26, p. 603–615.
K
LUTH
, C.F.,
AND
C
ONEY
, P.J., 1981, Plate tectonics of the Ancestral Rocky Mountains:
Geology, v. 9, p. 10–15.
K
ORTE
, C., J
ASPER
, T., K
OZUR
, H.W.,
AND
V
EIZER
,J.,2006,
87
Sr/
86
Sr record of Permian
seawater: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 240, p. 89–107.
K
OZUR
, H.W.,
AND
W
EEMS
, R.E., 2010, The biostratigraphic importance of conchos-
tracans in the continental Triassic of the northern hemisphere, in Lucas, S.G., ed., The
Triassic Timescale: Geological Society of London, Special Publication 334, no. 1,
p. 315–417.
K
UTZBACH
, J.E.,
AND
G
ALLIMORE
,R.G.,1989,Pangaeanclimates:megamonsoonsofthe
megacontinent: Journal of Geophysical Research, v. 94, p. 3341–3357.
L
AST
,W.M.,1990,Lacustrinedolomite:anoverviewofmodern,Holocene,and
Pleistocene occurrences: Earth-Science Reviews, v. 27, p. 221–263.
L
OWENSTEIN
, T., L
I
, J., B
ROWN
, C., R
OBERTS
, S., K
U
, T., L
UO
, S.,
AND
Y
ANG
,W.,1999,
200 ky paleoclimate record from Death Valley salt core: Geology, v. 27, p. 3–6.
L
OWENSTEIN
, T., H
EIN
, M., B
OBST
, A., J
ORDAN
, T., K
U
, T.,
AND
L
UO
,S.,2003,An
assessment of stratigraphic completeness in climate-sensitive closed-basin lake
sediments: Salar de Atacama, Chile: Journal of Sedimentary Research, v. 73, p. 91–104.
L
UZA
,K.V.,1978,RegionalseismicandgeologicevaluationsofNemahaUplift,
Oklahoma, Kansas, and Nebraska: Oklahoma Geology Notes, v. 38, p. 49–58.
M
ACHEL
, H.G., 2004, Concepts and models of dolomitization: a critical reappraisal, in
Braithwaite, C.J.R., Rizzi, G., and Darke, G., eds., The Geometry and Petrogenesis of
Dolomite Hydrocarbon Reservoirs: Geological Society of London, Special Publica-
tion 235, p. 7–63.
M
ARTENS
, U., W
EBER
, B.,
AND
V
ALENCIA
, V.A., 2010, U/Pb geochronology of Devonian
and older Paleozoic beds in the southeastern Maya block, Central America: Its
affinity with peri-Gondwanan terranes: Geological Society of America, Bulletin,
v. 122, p. 815–829.
M
AZZULLO
,S.J.,1999,EustaticandtectoniccontrolsoncyclicdepositioninLower
Permian ramp facies (Chase Group and basal Wellington formation) in the U.S.
Midcontinent, in Harris, P.M., Saller, A.H., and Simo, J.A., eds., Advances in
Carbonate Sequence Stratigraphy: Application to Reservoirs, Outcrops, and Models:
SEPM, Special Publication 63, p. 151–168.
M
C
K
EE
, E.D.,
AND
O
RIEL
,S.S.,1967,PaleotectonicInvestigationsofthePermian
System in the United States: United States Geological Survey, Professional Paper 515,
271 p.
M
C
L
ENNAN
,S.M.,H
EMMING
,S.,M
C
D
ANIEL
,D.K.,
AND
H
ANSON
, G.N., 1993,
Geochemical approaches to sedimentation, provenance, and tectonics, in Johnsson,
M.J., and Basu, A., eds., Processes Controlling the Composition of Clastic Sediments:
Geological Society of America, Special Paper 284, p. 21–40.
M
ONTENAT
, C., B
ARRIER
, P., O
TT D’
E
STEVOU
, P.,
AND
H
IBSCH
, C., 2007, Seismites: an
attempt at critical analysis and classification: Sedimentary Geology, v. 196, p. 1–4.
N
ETTLETON
, W.D.,
AND
S
LEEMAN
,J.R.,1985,Micromorphologyofvertisols,inDouglas,
L.A., and Thompson, M.L., eds., Soil Micromorphology and Soil Classification:
Madison, Soil Science of America, Special Publication 15, p. 165–196.
N
ICHOLSON
,S.E.,1999,HistoricalandmodernfluctuationsofLakesTanganyikaand
Rukwa and their relationship to rainfall variability: Climate Change, v. 41, p. 53–71.
N
URY
, D.,
AND
S
CHREIBER
, B.C., 1997, The Paleogene basins of southern Provence, in
Busson, G., and Schreiber, B.C., eds., Sedimentary Deposition in Rift and Foreland
Basins in France and Spain (Paleogene and Lower Neogene): New York, Columbia
University Press, p. 240–300.
O’C
ONNOR
, H., Z
ELLER
, D., B
AYNE
, C., J
EWETT
, J.,
AND
S
WINEFORD
,A.,1968,Permian
System, in Zeller, D., ed., The Stratigraphic Succession in Kansas: Kansas Geological
Survey, Bulletin 189, p. 43–53.
O
LSEN
,P.,1986,A40-million-yearlakerecordofEarlyMesozoicorbitalclimatic
forcing: Science, v. 234, p. 842–848.
P
ACK
, J.M., 2010, Paleoclimatic Implications of the Depositional Setting and Origin of
Sediments in the Wellington Formation, Kay County, Oklahoma [unpublished M.S.
thesis]: University of Oklahoma, Norman, Oklahoma, 192 p.
P
ARK
, L.E.,
AND
G
IERLOWSKI
-K
ORDESCH
, E.H., 2007, Paleozoic lake faunas: establishing
aquatic life on land: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 249,
p. 160–179.
P
ARRISH
, J.T., 1993, Climate of the supercontinent of Pangea: Journal of Geology,
v. 101, p. 215–233.
P
OPPE
,L.J.,P
ASKEVICH
,V.F.,H
ATHAWAY
,J.C.,
AND
B
LACKWOOD
, D.S., 2002,
A Laboratory Manual for X-Ray Powder Diffraction: U.S. Geological Survey,
Open-File Report 1-41, 88 p.
844 J.M. GILES ET AL.
JSR
P
RICE
, J., H
OGAN
, J.,
AND
G
ILBERT
, M., 1996, Rapakivi texture in the Mount
Scott granite, Wichita mountains, Oklahoma: European Journal of Mineralogy,
v. 8, p. 435.
P
YE
, K., 1987, Aeolian Dust and Dust Deposits: London, Academic Press Inc., 334 p.
Q
IANG
, M., C
HEN
, F., Z
HANG
, J., Z
U
, R., J
IN
, M., Z
HOU
, A.,
AND
X
IAO
, S., 2007, Grain
size in sediments from Lake Sugan: a possible linkage to dust storm events at the
northern margin of the Qinghai–Tibetan Plateau: Environmental Geology, v. 51,
p. 1229–1238.
Q
IAO
, Y., Z
HAO
, Z., W
ANG
, Y., F
U
, J., W
ANG
, S.,
AND
J
IANG
, F., 2009, Variations of
geochemical compositions and the paleoclimatic significance of a loess–soil sequence
from GarzeˆCountyofwesternSichuanProvince,China:ChineseScienceBulletin,
v. 54, p. 4697–4703.
R
AASCH
, G.O., 1946, The Wellington Formation in Oklahoma [unpublished Ph.D.
thesis]: University of Wisconsin, Madison, 157 p.
R
EHEIS
, M.C., G
OODMACHER
, J.C., H
ARDEN
, J.W., M
C
F
ADDEN
,L.D.,R
OCKWELL
, T.K.,
S
HROBA
, R.R., S
OWERS
, J.M.,
AND
T
AYLOR
, E.M., 1995, Quaternary soils and dust
deposition in southern Nevada and California: Geological Society of America,
Bulletin, v. 107, p. 1003–1022.
R
ETALLACK
, G.J., 1990, Soils of the Past; An Introduction to Paleopedology: Boston,
Unwin Hyman, 520 p.
S
A
´NCHEZ
-R
OMA
´N
, M., V
ASCONCELOS
, C., S
CHMID
, T., D
ITTRICH
, M., M
C
K
ENZIE
, J.A.,
Z
ENOBI
, R.,
AND
R
IVADENEYRA
,M.A.,2008,Aerobicmicrobialdolomiteatthe
nanometer scale: implications for the geologic record: Geology, v. 36, p. 879–882.
S
A
´NCHEZ
-R
OMA
´N
, M., M
C
K
ENZIE
, J.A.,
DE
L
UCA
R
EBELLO
W
AGENER
, A., R
IVADENEYRA
,
M.A.,
AND
V
ASCONCELOS
, C., 2009, Presence of sulfate does not inhibit low-temperature
dolomite precipitation: Earth and Planetary Science Letters, v. 285, p. 131–139.
S
ANZ
-M
ONTERO
, M.E., R
ODRIGUEZ
-A
RANDA
, J.P.,
AND
P
EREZ
-S
OBA
, C., 2009, Microbial
weathering of Fe-rich phyllosilicates and formation of pyrite in the dolomite
precipitating environment of a Miocene lacustrine system: European Journal of
Mineralogy, v. 21, p. 163–175.
S
CHREIBER
, B.C., 1978, Environments of Subaqueous Gypsum Deposition: SEPM, Short
Course 4, p. 43–73.
S
CHULTZ
, D.J., 1975, Crystalline Silica in Mudrocks [unpublished M.S. thesis]:
University of Oklahoma, Norman, Oklahoma, 48 p.
S
COTESE
, C.R., B
AMBACH
, R.K., B
ARTON
, C., V
OO
, R.V.D.,
AND
Z
IEGLER
,A.M.,1979,
Paleozoic base maps: Journal of Geology, v. 87, p. 217–277.
S
MALLEY
, I., K
UMAR
, R., O’H
ARA
D
HAND
, K., J
EFFERSON
, I.,
AND
E
VANS
, R., 2005, The
formation of silt material for terrestrial sediments: Particularly loess and dust:
Sedimentary Geology, v. 179, p. 321–328.
S
MITH
, G.,
AND
B
ISCHOFF
, J., 1997, An 800,000-Year Paleoclimatic Record from Core
OL-92, Owens Lake, Southeast California: Geological Society of America, Special
Paper 317, 165 p.
S
MITH
, G., B
ARCZAK
, V., M
OULTON
, G.,
AND
L
IDDICOAT
,J.,1983,CoreKM-3,aSurface-
to-Bedrock Record of Late Cenozoic Sedimentation in Searles Valley, California: U.S.
Geological Survey, Professional Paper 1256, 24 p.
S
MOOT
, J.P., 1993, Field trip guide: Quaternary–Holocene–Lacustrine Sediments of
Lake Lahontan, Truckee River Canyon North of Wadsworth, Nevada: U.S.
Geological Survey, Open-File Report 93-0689, 35 p.
S
MOOT
, J.P.,
AND
C
ASTENS
-S
EIDELL
, B., 1994, Sedimentary features produced by
efflorescent salt crusts, Saline Valley and Death Valley, California, in Renaut,
R.W., and Last, W.M., eds., Sedimentology and Geochemistry of Modern and
Ancient Saline Lakes: SEPM, Special Publication 50, p. 73–90.
S
MOOT
, J.P.,
AND
L
OWENSTEIN
, T.K., 1991, Depositional environments of non-marine
evaporites, in Melvin, J., ed., Evaporites, Petroleum, and Mineral Resources:
Amsterdam, Elsevier, Developments in Sedimentology 50, p. 189–347.
S
OREGHAN
, G.S.,
AND
S
OREGHAN
, M.J., 2013, Tracing clastic delivery to the Permian
Delaware Basin, U.S.A.: implications for paleogeography and circulation in
westernmost equatorial Pangea: Journal of Sedimentary Research, v. 83, p. 786–802.
S
OREGHAN
, G., E
LMORE
, R.,
AND
L
EWCHUK
, M., 2002, Sedimentologic–magnetic record
of western Pangean climate in upper Paleozoic loessite (lower Cutler beds, Utah):
Geological Society of America, Bulletin, v. 114, p. 1019–1035.
S
OREGHAN
, G.S., K
ELLER
, G.R., G
ILBERT
, M.C., C
HASE
, C.G.,
AND
S
WEET
, D.E., 2012,
Load-induced subsidence of the Ancestral Rocky Mountains recorded by preservation
of Permian landscapes: Geosphere 8, p. 654–668.
S
OREGHAN
, M.J.,
AND
S
OREGHAN
,G.S.,2007,Whole-rockgeochemistryofupper
Paleozoic loessite, western Pangaea: implications for paleo-atmospheric circulation:
Earth and Planetary Science Letters, v. 255, p. 117–132.
S
OREGHAN
, M., S
OREGHAN
, G.,
AND
H
AMILTON
, M., 2002, Paleowinds inferred from
detrital-zircon geochronology of upper Paleozoic loessite, western equatorial Pangea:
Geology, v. 30, p. 695–698.
S
OREGHAN
, M.J., S
OREGHAN
, G.S.,
AND
H
AMILTON
, M.A., 2008, Glacial–interglacial
shifts in atmospheric circulation of western tropical Pangaea: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 268, p. 260–272.
S
OYINKA
, O.A.,
AND
S
LATT
, R.M., 2008, Identification and micro-stratigraphy of
hyperpycnites and turbidites in Cretaceous Lewis Shale, Wyoming: Sedimentology,
v. 55, p. 1117–1133.
S
PENCER
, R.J.,
AND
L
OWENSTEIN
,T.K.,1990,Evaporites:GeoscienceCanada,Reprint
Series, v. 4, p. 141–163.
S
UTTON
, S.J.,
AND
L
AND
, L.S., 1996, Postdepositional chemical alteration of Ouachita
shales: Geological Society of America, Bulletin, v. 108, p. 978–991.
S
WANSON
, B.A.,
AND
C
ARLSON
, K.J., 2002, Walk, wade, or swim? Vertebrate traces on an
Early Permian lakeshore: Palaios, v. 17, p. 123–133.
S
WEET
, D.,
AND
S
OREGHAN
, G., 2010, Late Paleozoic tectonics and paleogeography of the
ancestral Front Range: structural, stratigraphic, and sedimentologic evidence from the
Fountain Formation (Manitou Springs, Colorado): Geological Society of America,
Bulletin, v. 122, p. 575–594.
S
WINEFORD
, A.,
AND
R
UNNELS
, R.T., 1953, Identification of polyhalite (a potash mineral)
in Kansas Permian salt: Kansas Academy of Science Transactions, v. 56, p. 364–
370.
T
ABOR
, N.,
AND
M
ONTAN
˜EZ
, I., 2002, Shifts in late Paleozoic atmospheric circulation
over western equatorial Pangea: insights from pedogenic mineral d
18
Ocompositions:
Geology, v. 30, p. 1127–1130.
T
ABOR
, N.J.A.,
AND
P
OULSEN
,C.J.,2008,PalaeoclimateacrosstheLatePennsylvanian
Early Permian tropic al palaeolatitudes: a review of climate indicators, their
distribution, and relation to palaeophysiographic climate factors: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 268, p. 293–310.
T
ABOR
,N.J.,M
ONTAN
˜EZ
,I.P.,S
COTESE
,C.R.,P
OULSEN
,C.J.,
AND
,M
ACK
,G.H,2008,
Paleosol archives of environmental and climatic history in paleotropical western
Pangea during the latest Pennsylvanian through Early Permian, in Fielding,
C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age
in Time and Space: Geological S ociety of America, Special Publication 441, p. 291–
303.
T
ASCH
, P., 1958, Permian Conchostracan-Bearing Beds of Kansas: Part 1. Jester Creek
Section: Fauna and Paleoecology: Journal of Paleontology, v. 32, p. 525–540.
T
ASCH
,P.,1961a,DataonsomenewLeonardianconchostracanswithobservationson
the taxonomy of the family vertexiidae: Journal of Paleontology, v. 35, p. 1121–1129.
T
ASCH
,P.,1961b,Paleolimnology:Part2,HarveyandWedgwickCounties,Kansas,
stratigraphy and biota: Journal of Paleontology, v. 35, p. 836–865.
T
ASCH
,P.,1962,Taxonomicandevolutionarysignificanceoftwonewconchostracan
genera from the Midcontinent Wellington Formation: Journal of Paleontology, v. 36,
p. 817–821.
T
ASCH
,P.,1964,PeriodicityintheWellingtonFormationofKansasandOklahoma:
Kansas Geological Survey, Bulletin 169, p. 481–496.
T
AYLOR
, S.R.,
AND
M
C
C
LENNAN
,S.M.,1985,TheContinentalCrust;ItsComposition
and Evolution; An Examination of the Geochemical Record Preserved in Sedimentary
Rocks: Oxford, Blackwell Scientific Publications, Geoscience Texts, 312 p.
T
AYLOR
, S.R.,
AND
M
C
L
ENNAN
,S.M.,2001,Chemicalcompositionandelement
distribution in the Earth’s crust, in Robert, A.M., ed., Encyclopedia of Physical
Science and Technology: New York, Academic Press, p. 697–719.
T
OTTEN
,M.W.,H
ANAN
,M.A.,
AND
W
EAVER
, B.L., 2000, Beyond whole-rock
geochemistry of shales: the importance of assessing mineralogic controls for revealing
tectonic discriminants of multiple sediment sources for the Ouachita Mountain flysch
deposits: Geological Society of America, Bulletin, v. 112, p. 1012–1022.
T
RUDINGER
,P.A.,C
HAMBERS
,L.A.,
AND
S
MITH
,J.W.,1985,Low-temperaturesulphatereduction:
biological versus abiological: Canadian Journal of Earth Sciences, v. 22, p. 1910–1918.
T
SOAR
, H.,
AND
P
YE
,K.,1987,Dusttransportandthequestionofdesertloess
formation: Sedimentology, v. 34, p. 139–153.
T
WENHOFEL
,W.,1926,GeologyoftheMinganIslands:GeologicalSocietyofAmerica,
Bulletin, v. 37, p. 535–550.
V
ANDERVOORT
, D.S., 1997, Stratigraphic response to saline lake-level fluctuations and
the origin of cyclic nonmarine evaporite deposits: The Pleistocene Blanca Lila
Formation, northwest Argentina: Geological Society of America, Bulletin, v. 109,
p. 210–224.
V
AN
S
CHMUS
, W., B
ICKFORD
, M., A
NDERSON
, J., B
ENDER
, E., A
NDERSON
, R., B
AUER
, P.,
R
OBERTSON
, J., B
OWRING
, S., C
ONDIE
, K.,
AND
D
ENISON
, R., 1993, Transcontinental
Proterozoic provinces, Precambrian, The Geological Society of America, Geology of
North America, v. C-3, Conterminous US, p. 171–334.
V
ASCONCELOS
, C.,
AND
,M
C
K
ENZIE
, J.A., 1997, Microbial mediation of modern dolomite
precipitation and diagenesis under anoxic conditions (Lagoa Vermelha, Rio de
Janerio, Brazil); Journal of Sedimentary Research v. 67, p. 378–390.
V
EEVERS
, J.J.,
AND
P
OWELL
, M., 1987, Late Paleozoic glacial episodes in Gondwanaland
reflected in transgressive–regressive depositional sequences in Euramerica: Geological
Society of America, Bulletin, v. 98, p. 475–487.
V
ER
W
IEBE
, W.A., 1937, The Wellington Formation of central Kansas: Wichita
University, Bulletin 12, p. 3–18.
VON DER
B
ORCH
, C.C., 1976, Stratigraphy and formation of Holocene dolomitic
carbonate deposits of the Coorong area, South Australia: Journal of Sedimentary
Petrology, v. 46, p. 952–966.
W
ANLESS
, H.R.,
AND
S
HEPARD
,F.P.,1936,Sealevelandclimaticchangesrelatedtolate
Paleozoic cycles: Geological Society of America, Bulletin, v. 47, p. 1177–1206.
W
ARREN
, J.K., 1989, Evaporite Sedimentology: Importance in Hydrocarbon Accumu-
lation: Englewood Cliffs, New Jersey, Prentice-Hall, 285 p.
W
ARREN
, J.K.,
AND
K
ENDALL
, C.G.S
T
.C., 1985, Comparison of sequences formed in
marine sabkha (subaerial) and salina (subaqueous) settings; modern and ancient:
American Association of Petroleum Geologists, Bulletin, v. 69, p. 1013–1023.
W
ATNEY
,W.L.,B
ERG
, J.,
AND
P
AUL
,S.,1988,OriginanddistributionoftheHutchinson
Salt Member (lower Leonardian) in Kansas, in Morgan, W.A., and Babcock, J.A.,
eds., Permian Rocks of the Mid-Continent: Midcontinent SEPM, Special Publication
1, p. 113–135.
W
EBER
, B., S
CHAAF
, P., V
ALENCIA
, V.A., I
RIONDO
, A.,
AND
O
RTEGA
-G
UTIE
´RREZ
, F., 2006,
Provenance ages of late Paleozoic sandstones (Santa Rosa Formation) from the Maya
Block, SE Mexico: implications on the tectonic evolution of western Pangea: Revista
Mexicana de Ciencias Geolo´gicas, v. 23, p. 262–276.
LAKES, LOESS, AND PALEOSOLS IN THE PERMIAN WELLINGTON FORMATION OF OKLAHOMA 845
JSR
W
ILLIAMS
-S
TROUD
, S.C., 1994, The evolution of an inland sea of marine origin to a non-
marine saline lake: the Pennsylvanian Paradox salt, in Renaut, R.W., and Last, W.M.,
eds., Sedimentology and Geochemistry of Modern and Ancient Saline Lakes: SEPM,
Special Publication 50, p. 293–306.
W
ILSON
, B., 1927, Paleogeography of the upper Paleozoic rocks of Oklahoma:
Oklahoma Geological Survey, Bulletin 41, p. 22–86.
W
ILSON
, L., 1962, Permian plant microfossils from the Flowerpot Formation, Greer
County, Oklahoma: Oklahoma State Geological Survey, Circular 49, 50 p.
W
RIGHT
, V.P.,
AND
M
ARRIOTT
, S.B., 2007, The dangers of taking mud for granted:
lessons from Lower Old Red Sandstone dryland river systems of South Wales:
Sedimentary Geology, v. 195, p. 91–100.
X
IAO
,J.,A
N
,Z.,L
IU
,T.,I
NOUCHI
,Y.,K
UMAI
,H.,Y
OSHIKAWA
,S.,
AND
K
ONDO
,Y.,1999,East
Asian monsoon variation during the last130,000 years: evidence from the LoessPlateau of
central China and Lake Biwa of Japan: Quaternary Science Reviews, v. 18, p. 147–157.
Y
ANG
,S.V.,1985,PetrologicalandGeochemicalApproachestotheOriginoftheSan
Angelo–Flowerpot Red Beds (Permian) and Their Associated Stratiform Copper
Mineralizations in North Central Texas and Southwestern Oklahoma [unpublished
Ph.D. thesis]: University of Texas at Dallas, 318 p.
Y
ECHIELI
, Y.,
AND
W
OOD
, W.W., 2002, Hydrogeologic processes in saline systems:
playas, sabkhas, and saline lakes: Earth-Science Reviews, v. 58, p. 343–365.
Received 4 January 2012; accepted 24 April 2012.
846 J.M. GILES ET AL.
JSR
... An absence of alternative explanations for delivery of the sediment (e.g., absence of fluvial features) strengthens an interpretation of eolian delivery. Loessite has also been identified in core, based primarily on documentation of massive (unstructured), monotonous siltstone (e.g., Kessler et al., 2001;Dubois et al., 2012;Giles et al., 2013), and random grain-fabric orientations (Wilkins et al., 2018). Sedimentary structures (e.g., ripples, desiccation cracks, laminations) are common in silt-rich strata otherwise posited to record eolian delivery, indicating reworking by water or pedogenic processes (Johnson, 1989; Soreghan et al., 2008b;Giles et al., 2013;Sweet et al., 2013;Foster et al., 2014;Wilkins et al., 2018;Pfeifer et al., 2020), since wet surfaces enhance dust trapping. ...
... Loessite has also been identified in core, based primarily on documentation of massive (unstructured), monotonous siltstone (e.g., Kessler et al., 2001;Dubois et al., 2012;Giles et al., 2013), and random grain-fabric orientations (Wilkins et al., 2018). Sedimentary structures (e.g., ripples, desiccation cracks, laminations) are common in silt-rich strata otherwise posited to record eolian delivery, indicating reworking by water or pedogenic processes (Johnson, 1989; Soreghan et al., 2008b;Giles et al., 2013;Sweet et al., 2013;Foster et al., 2014;Wilkins et al., 2018;Pfeifer et al., 2020), since wet surfaces enhance dust trapping. ...
... 16 paleoprecipitation; Maher, 2016). Until relatively recently, loess was recognized exclusively from the Quaternary record, and linked in part to the widespread icehouse conditions of the Quaternary; but now paleoloess has been recognized in Earth's deep-time record, especially from the Carboniferous-Permian icehouse (Murphy, 1987;Johnson, 1989;Kessler et al., 2001;Mack and Dinterman, 2002;Soreghan et al., 2002;Tramp et al., 2004;Soreghan et al., 2007;Soreghan et al., 2008a;Giles et al., 2013;Sweet et al., 2013;Foster et al., 2014;Pfeifer et al., 2020), and the Neoproterozoic icehouse (Edwards, 1979;Retallack, 2011;Retallack et al., 2015). But a few examples of inferred Mesozoic (mostly Triassic; one Cretaceous) loessite (Chan, 1999;Jefferson et al., 2002;Lawton et al., 2018;Wilkins et al., 2018;Wilson et al., 2020) exist, deposited during greenhouse intervals, which calls into question what if any attributes of loess might specifically indicate a glacial genesis. ...
Article
Full-text available
Earth has sustained continental glaciation several times in its past. Because continental glaciers ground to low elevations, sedimentary records of ice contact can be preserved from regions that were below base level, or subject to subsidence. In such regions, glaciated pavements, ice-contact deposits such as glacial till with striated clasts, and glaciolacustrine or glaciomarine strata with dropstones reveal clear signs of former glaciation. But assessing upland (mountain) glaciation poses particular challenges because elevated regions typically erode, and thus have extraordinarily poor preservation potential. Here we propose approaches for detecting the former presence of glaciation in the absence or near-absence of ice-contact indicators; we apply this specifically to the problem of detecting upland glaciation, and consider the implications for Earth’s climate system. Where even piedmont regions are eroded, pro- and periglacial phenomena will constitute the primary record of upland glaciation. Striations on large (pebble and larger) clasts survive only a few km of fluvial transport, but microtextures developed on quartz sand survive longer distances of transport, and record high-stress fractures consistent with glaciation. Proglacial fluvial systems can be difficult to distinguish from non-glacial systems, but a preponderance of facies signaling abundant water and sediment, such as hyperconcentrated flood flows, non-cohesive fine-grained debris flows, and/or large-scale and coarse-grained cross-stratification are consistent with proglacial conditions, especially in combination with evidence for cold temperatures, such as rip-up clasts composed of noncohesive sediment, indicating frozen conditions, and/or evidence for a predominance of physical over chemical weathering. Other indicators of freezing (periglacial) conditions include frozen-ground phenomena such as fossil ice wedges and ice crystals. Voluminous loess deposits and eolian-marine silt/mudstone characterized by silt modes, a significant proportion of primary silicate minerals, and a provenance from non-silt precursors can indicate the operation of glacial grinding, even though such deposits may be far removed from the site(s) of glaciation. Ultimately, in the absence of unambiguous ice-contact indicators, inferences of glaciation must be grounded on an array of observations that together record abundant meltwater, temperatures capable of sustaining glaciation, and glacial weathering (e.g., glacial grinding). If such arguments are viable, they can bolster the accuracy of past climate models, and guide climate modelers in assessing the types of forcings that could enable glaciation at elevation, as well as the extent to which (extensive) upland glaciation might have influenced global climate.
... Enhancing our understanding of the paleogeography and depositional processes of this broader region could have significant implications beyond oil and gas exploration. It would be particularly relevant for paleoclimate studies (Benison and Goldstein, 2000;Giles et al., 2013;Soreghan and Soreghan, 2013), sequence stratigraphy (Thompson et al., 2018), carbon storage research (Ren and Duncan, 2019), and also for groundwater monitoring and aquifer mapping (Gardiner et al., 2020). ...
Article
Altered volcanic ash, preserved as bentonite clay within sedimentary depositional units, can facilitate regional stratigraphic correlation. While many bentonite ash beds have been documented in core extracted over decades of hydrocarbon exploration in the Delaware Basin, a particular ~1 cm-thick bentonite has been identified at a distinct stratigraphic position, at a sedimentary facies transition from thin-bedded calciclastic turbidites to thick�bedded siliciclastic turbidites, in the compositionally-mixed upper Wolfcamp formation. A comparison of geochemical analyses of the ash from two cores and petrophysical correlations across eight wells was conducted to test the regional correlation of the ash. Despite diagenetic alteration since seafloor deposition ~280 Ma, nearly identical geochemical trends between incompatible elemental subsets, including light and heavy rare earth elements, large ion lithophiles, and high field strength elements, suggest that these two ash deposits originated from the same volcanic center. Detailed sedimentologic analyses and lithostratigraphic correlations across three cores spaced tens of km apart confirm that the ash bed is bound by correlative quiescent deepwater strata. Based on the chemical composition, probable timing of deposition, and proximity, we speculate that the Las Delicias Arc produced this ash, and it was transported by wind and deposited as a hemipelagite into the Delaware Basin deep marine environment. This discovery implies the existence of the ash marker in other cores within Late Paleozoic western equatorial Pangea, e.g. in compositionally-mixed deep marine deposits of the Midland Basin and outcropping deposits around the Greater Permian Basin and beyond, which would facilitate regional chronostratigraphic correlation for a critical time in Earth’s history.
... Enhancing our understanding of the paleogeography and depositional processes of this broader region could have significant implications beyond oil and gas exploration. It would be particularly relevant for paleoclimate studies (Benison and Goldstein, 2000;Giles et al., 2013;Soreghan and Soreghan, 2013), sequence stratigraphy (Thompson et al., 2018), carbon storage research (Ren and Duncan, 2019), and also for groundwater monitoring and aquifer mapping (Gardiner et al., 2020). ...
... There was also an upward increase in landscape stability and nutrient content as recorded by changes in pedotypes, paleosol thicknesses, and ichnofossil assemblages (Kraus 1999;Hasiotis 2007). This is similar to trends observed in contemporaneous localities in Kansas, Oklahoma, and Texas (Kessler et al. 2001;Tabor and Montañez 2004;DiMichele et al. 2006;Mack et al. 2010;Giles et al. 2013;Counts and Hasiotis 2014;Tanner and Lucas 2017). The transition in pedotypes was not gradual in the study area, but was characterized by repeated changes in estimated paleoprecipitation from the chemical index of alteration and depth to carbonate horizons from~200 to 1400 mm/yr from the base of the Monongahela Group to the Dunkard Group (Hembree and Bowen 2017;Hembree and McFadden 2020). ...
Article
Full-text available
The late Paleozoic transition is well represented by the upper Pennsylvanian to lower Permian Conemaugh, Monongahela, and Dunkard groups of the western Appalachian Basin (U.S.A.). These units contain abundant paleosols possessing suites of ichnofossils that serve as indicators of soil moisture, soil organic content, water table level, precipitation, and landscape stability. Analysis of these units can, therefore, be used to refine the details of how late Paleozoic terrestrial landscapes changed through time. A study along a 50 km west-east and a 40 km north-south transect through southeast Ohio and southwest West Virginia resulted in the recognition of 24 pedotypes with distinct ichnofossil assemblages. Ichnofossils include rhizoliths, Planolites, Palaeophycus, Taenidium, Scoyenia, Macanopsis, Skolithos, Cylindricum, cf. Psilonichnus, Arenicolites, mottles, and coprolites produced by various plants, gastropods, and larval-to-adult soil arthropods. Soil-forming environments include palustrine, levee, proximal to distal floodplain, interfluve, backswamp, marsh, and fen settings. An up-section shift in pedotypes from Argillisols to Vertisols and Calcisols as well as an overall increase in the diversity of pedotypes recorded a change in soil-forming conditions, resulting in a diverse landscape that changed significantly as mean annual precipitation rose and fell. An up-section increase in ichnofossil diversity in the paleosols and changes in ichnocoenoses suggests an increased dependence on the soil as a refuge and as a food resource. Overall, growing instability of the climate during the Pennsylvanian–Permian transition led to a more heterogeneous landscape that helped to promote colonization of a more diverse assemblage of soil organisms.
... There was also an upward increase in landscape stability and nutrient content as recorded by changes in pedotypes, paleosol thicknesses, and ichnofossil assemblages (Kraus 1999;Hasiotis 2007). This is similar to trends observed in contemporaneous localities in Kansas, Oklahoma, and Texas (Kessler et al. 2001;Tabor and Montañez 2004;DiMichele et al. 2006;Mack et al. 2010;Giles et al. 2013;Counts and Hasiotis 2014;Tanner and Lucas 2017). The transition in pedotypes was not gradual in the study area, but was characterized by repeated changes in estimated paleoprecipitation from the chemical index of alteration and depth to carbonate horizons from~200 to 1400 mm/yr from the base of the Monongahela Group to the Dunkard Group (Hembree and Bowen 2017;Hembree and McFadden 2020). ...
... Enhancing our understanding of the paleogeography and depositional processes of this broader region could have significant implications beyond oil and gas exploration. It would be particularly relevant for paleoclimate studies (Benison and Goldstein, 2000;Giles et al., 2013;Soreghan and Soreghan, 2013), sequence stratigraphy (Thompson et al., 2018), carbon storage research (Ren and Duncan, 2019), and also for groundwater monitoring and aquifer mapping (Gardiner et al., 2020). ...
... Aeolian sediments are widespread in geological history (Muhs, 2007). In addition to the well-studied Cenozoic aeolian loess worldwide Meijer et al., 2020), for example, Paleozoic, Permian and Triassic aeolian loess or loessite, have also been identified and investigated for their palaeoenvironmental implications (Giles et al., 2013;Soreghan et al., 2008;Wilkins et al., 2018;Wilson et al., 2020). The aeolian loess origin of the Mercia mudstone was primarily proposed based on its collapse behaviour when wetted or loaded (Bosworth, 1913;Jefferson et al., 2000), analogous to the Quaternary red clay of Australia (parna by some authors) in comparable palaeoenvironments (Talbot et al., 1994). ...
Article
Late Permian and Triassic red mudstones are widely distributed and exposed in southwestern England. This homogenous, massive mudstone succession has been extensively studied for its palaeoenvironmental potential. However, its origin and palaeoenvironment remain under debate. The Triassic Mercia mudstone was early proposed to be aeolian loessic clay based on collapse nature when wetted or loaded and its similarities with aeolian clay in Quaternary Australia, but lacks experimental confirmation and is not generally accepted. In comparison with aeolian red clay in China, here we confirmed that the Triassic Mercia mudstone, together with the Upper Permian Aylesbeare mudstone, was dominated by aeolian clay. This was based on lines of evidence regarding sedimentary structure, particle grain size, geochemistry, micromorphology, and pedogenic features. The results showed that the Mercia mudstone and the Aylesbeare mudstone had sedimentary structures and grain size distributions that were similar to the aeolian red clay in the Chinese Loess Plateau. The major chemical elements correlated well with the aeolian red clay, with the exception of minor discrepancies in CaO and K2O. The similarity in the rare earth element (REE) distribution of the Mercia mudstone and the Aylesbeare mustone with aeolian red clay inferred their similar origin. Widespread pedogenic features indicated the universal development of palaeovertisols forming in alternating wet/dry climates. It was hypothesised that Permo-Triassic southwestern England was dominated by aeolian dust accumulation in a seasonal wet/dry climate, whereas perennial lakes or longstanding rivers across the study area might not exist.
Article
In the last decade, the Fossilized Birth–Death (FBD) process has yielded interesting clues about the evolution of biodiversity through time. To facilitate such studies, we extend our method to compute the probability density of phylogenetic trees of extant and extinct taxa in which the only temporal information is provided by the fossil ages (i.e. without the divergence times) in order to deal with the piecewise constant FBD process, known as the “skyline FBD”, which allows rates to change between pre‐defined time intervals, as well as modelling extinction events at the bounds of these intervals. We develop approaches based on this method to assess hypotheses about the diversification process and to answer questions such as “Does a mass extinction occur at this time?” or “Is there a change in the fossilization rate between two given periods?”. Our software can also yield Bayesian and maximum‐likelihood estimates of the parameters of the skyline FBD model under various constraints. These approaches are applied to a simulated dataset in order to test their ability to answer the questions above. Finally, we study an updated dataset of Permo‐Carboniferous synapsids to get additional insights into the dynamics of biodiversity change in three clades (Ophiacodontidae, Edaphosauridae and Sphenacodontidae) in the Pennsylvanian (Late Carboniferous) and Cisuralian (Early Permian), and to assess support for end‐Sakmarian (or Artinskian) and end‐Cisuralian mass extinction events discussed in previous studies.
Article
Full-text available
Recognizing lungfish (Sarcopterygii, Dipnoi) estivation burrows: ichnotaxonomy and paleoenvironment. Lungfishes first appeared on Earth around 350 million years ago in freshwater environments from Gondwana, since when they suffered harsh adaptations through the geological time, among which, the aestivation capability. Such a behavior keeps preserved in the geological record along the eras and the fossil burrows own a high potential for environmental inferences. From so much we developed this work aiming to (i) gather the records once published on this theme, (ii) to discuss the adopted standard by the authors for interpreting the burrows and their burrowers, and (iii) to discuss the paleoenvironmental significance of the lungfish burrows presenting a model of its probable occurrence in the landscape context. For such a purpose, we made the literature review searching for terms related to the theme, in different basis and pages of scientific journals. We have found 35 articles reporting new occurrences of lungfish burrows, besides uncertain or refuted records, according to ichnotaxonomical parameters. A large proportion of the papers inform about sedimentary facies and paleoenvironmental conditions. Essentially, the records occur in areas of shallow rivers and lakes of a semiarid to subtropical climate, with seasonal humidity variations, but also in coastal environments. Still, from the presented publications, we consider lungfish aestivation burrows to be safe paleoenvironmental indicators. Keywords: lungfish; aestivation; ichnofossils; ichnotaxonomy; floodplains; semiarid.
Article
Chemical analyses of 1371 fluid inclusions in 131 halite samples with marine 87Sr/86Sr values were used to reconstruct the strontium concentrations [Sr]SW of Phanerozoic and Neoproterozoic seawater. [Sr]SW varied seven-fold and oscillated twice between high- and low-Sr concentrations over the past 550 million years (Myr), in rhythm with Ca-rich and SO4-poor paleoseawater intervals and calcite-aragonite seas. Variations in the [Sr]/[Ca]SW ratio from fluid inclusions were not significant over the past ~270 Myr, and are within ±3 µmol/mmol of the modern [Sr]/[Ca]SW ratio of ~8.5 µmol/mmol. These results agree with the [Sr]/[Ca]SW ratios obtained from fossil corals, benthic foraminifera, brachiopods, belemnites, and rudists. [Sr]/[Ca]SW in the early and middle Paleozoic was ~ 2 times the modern [Sr]/[Ca]SW ratio. A major shift of the [Sr]/[Ca]SW ratio in the late Permian coincided with the initial rifting of the Pangean supercontinent. Seawater 87Sr/86Sr ratios plotted against 1/[Sr]SW show two distinct linear correlations: negative correlation from 515–252 Ma and positive correlation from 150–0 Ma, suggesting different controls on the global Sr cycle between these intervals. The negative correlation coincides with the long-term assembly of Pangea in the Paleozoic (~500–250 Ma). The positive correlation from 150–0 Ma parallels the break-up of Pangea and the decrease of mid-ocean ridge (MOR) hydrothermal fluid flux and subduction zone length in the Mesozoic and Cenozoic.
Article
The upper member of the 200 m (660 ft) thick Morgan Formation (Middle Pennsylvanian) consists of 5-25 m (16-82 ft) thick, very fine-grained quartz sandstone units that are interbedded repetitively with 0.5-11 m (1.6-36 ft) thick, oolitic, bioclastic, peloidal, and micritic carbonate units. Similar repetitive sequences occur widely in western North America. The quartz sandstone-carbonate cyclothems defined by this study have potential as targets for hydrocarbon exploration. Both eolian dune sandstones and dolomitized shelf carbonate strata are locally important reservoir rocks in the subsurface in parts of the western Overthrust belt in Utah and Wyoming. 84 references, 22 figures, 4 tables.
Chapter
The Anadarko Basin of the southern midcontinent, U.S.A., is the deepest cratonic basin in North America, with an existing in-place sedimentary section of 12–13 km. It has a more complex history than some other interior basins because it is both a cratonic and a foreland basin. It shares a common origin with several North American Paleozoic basins in having formed over a preexisting rift zone. This rift zone — the Southern Oklahoma Aulacogen — was unusual, having been formed athwart the suture between the southern 1.35–1.40 Ga Granite-Rhyolite terrane and the 1.2 Ga Texas era ton. The rift was active in the late Proterozoic-early Cambrian, but may have inherited an earlier Proterozoic grain. The Paleozoic Anadarko Basin was areally similar to present-day Michigan or Illinois, but was segmented and deepened in the Pennsylvanian during the Ouachita collision. It had a poorly recognized Permian phase which perhaps extended into the Mesozoic. The Anadarko Basin had a clear thermal subsidence phase in the early Paleozoic. The basin was reactivated in the Mississippian, and this reactivation presumably required plate margin interactions in some form. The basin was thrust-loaded in the Pennsylvanian. The Permian subsidence must also relate to 1) plate margin effects not well understood, and 2) compaction and diagenesis of the Pennsylvanian fill.
Chapter
Cyclic marine carbonate and marine to terrestrial siliciclastic strata in the upper Wolfcampian Chase Group and basal Wellington Formation were deposited on a low-relief, south-sloping ramp overlying the buried Nemaha ridge. The section comprises eight depositional sequences, each of which consists of two higher-frequency cyclothems of transgressive-regressive character. Each cyclothem includes a transgressive systems tract and ensuing highstand systems tract, and the maximum flooding surface of each sequence consistently occurs in the basal cyclothem of that sequence. Cyclothems are separated by relative lowstand systems tracts of dominantly marginal-marine facies, whereas sequence boundaries are either regionally extensive unconformities or lowstand systems tracts of paleosols. Thickness and facies-stacking patterns of the sequences are arranged so as to define three third-order cycles; component sequences and cyclothems are regarded as fourth- and fifth-order cycles, respectively. The regional occurrence of these cycles is interpreted to suggest glacio-eustatic forcing. The complex internal architecture of the sequences reflects interplay among forcing parameters and periodic syndepositional tectonism along the Nemaha ridge. Although there is coincidence between maxima of tectonic activity and degree of architectural complexity of sequences within each third-order cycle, the episodic nature of tectonism does not appear to have been consistent enough to have forced cyclicity in this section. Rather, syndepositional tectonism merely overprinted thickness variations of systems tracts within sequences on the fundamentally glacio-eustatically forced sequence stratigraphic architecture of this section.
Article
The Baza Basin (the eastern part of the Guadix-Baza Basin, in SE Spain) underwent a significant evaporitic sedimentation in non-marine settings during the Plio-Pleistocene. The largest of these settings developed in the central part of the basin as a shallow, saline lake system, where abundant gypsum deposits occur (Benamaurel and Galera Gypsum units). This paper studies these deposits from the cartographic, stratigraphic, sedimentologic, petrographic and isotopic points of view. All the gypsum has been preserved as primary enabling a good characterization of the environmental conditions of the original precipitates. Three interconnected subenvironments or zones -inner, intermediate and marginal- are differentiated in this saline lake system. In the inner zone (shallow to relatively deeper central lake), a regular alternation of gypsum and carbonate laminae accumulated suggesting a seasonal regime. Many of the gypsum laminae display size-gradation (both reverse and symmetrical) indicating free precipitation in a stable brine body. In the intermediate zone (mosaic of shallow lakes), carbonate (dolomitic) beds formed in association with dark lutite levels, gypsum beds, and gypsum nodules and micronodules. Moreover, anoxic conditions developed in this zone, which resulted in the formation of native sulphur occurrences of economic value. In the marginal zone, the evaporitic deposits are developed only locally (the Galera Gypsum unit, in particular). The various gypsum lithofacies in these marginal deposits suggest the presence of a number of shallow to exposed settings (saline mudflat, saline marsh and gypsiferous pond). The isotopic composition (delta S-34, delta O-18) of the gypsum samples is consistent with a Triassic origin of the sulphate as a result of chemical recycling. These data also suggest that a similar recycling mechanism operated in the saline lake system, where the gypsum sediments from the margins were coevally recycled towards the inner part. Both the evaporitic mineralogy and the gypsum and carbonate lithofacies suggest that semiarid conditions prevailed in this basin during the accumulation of the evaporite units. (c) 2007 Elsevier B.V. All rights reserved.
Article
Quantitative techniques employing Markovian processes have been used by a number of investigators to test stratigraphic sections against a null hypothesis of a random succession of sedimentary facies. Rejection of the null hypothesis is used to imply presence of a Markov chain in a succession. However, many of these techniques possess serious statistical flaws arising from the structuring of facies transitions. The result is that presence of preferred facies and regular 'cyclic' behavior in sedimentary successions cannot be rigorously supported. Log-linear models, in conjunction with a stepwise procedure for identification of significant transitions, provide a rigorous method for determining the presence and extent of the Markov property in stratigraphic sections. A wide variety of geologic problems that are structured in terms of temporal succession can be examined as a Markov process using log-linear models. -Author.
Article
Cores display characteristics other than the weathered-clay or caliche horizons or underclay type of paleosol profiles from the Upper Pennsylvanian in the Midcontinent. Paleosol features include clay-filled fractures and solution channels (cryptokarst) in the top of the Iatan Limestone which display some degree of secondary clay orientation along the edges (neostrians). Pedogenesis in this unit resulted from subaerial exposure during retreat of the sea following the transgressive-regressive marine cycle that deposited the Iatan Limestone. Solution-channelling of the top of the Iatan and leaching of clays within its fractures indicates initial exposure to a humid climate with strong meteoric flux. The Stranger paleosol, however, displays features of Entisol, Inceptisol, or Vertisol development in terrestrial deposits under a somewhat drier climate following the partial solution of the upper Iatan. -from Authors