ArticlePDF Available

The paleoclimatic and geochronologic utility of coring redbeds and evaporites: A case study from the RKB core (Permian, Kansas, USA)

Authors:

Abstract

Drill core is critical for robust and high-resolution reconstructions of Earth’s climate record, as well demonstrated from both marine successions and modern long-lived lake systems. Deep-time climate reconstructions increasingly require core-based data, but some facies, notably red beds and evaporites, have garnered less attention for both paleoclimatic and geochronologic analyses. Here, we highlight studies from the Rebecca K. Bounds (RKB) core, a nearly continuous, >1.6 km drill core extending from the Cretaceous to the Mississippian, recovered from the US Midcontinent by Amoco Production Company in 1988, and serendipitously made available for academic research. Recent research conducted on this core illustrates the potential to recover high-resolution data for geochronologic and climatic reconstructions from both the fine-grained red bed strata, which largely represent paleo-loess deposits, and associated evaporite strata. In this case, availability of core was instrumental for (1) accessing a continuous vertical section that establishes unambiguous superposition key to both magnetostratigraphic and paleoclimatic analyses, and (2) providing pristine sample material from friable, soluble, and/or lithofacies and mineralogical species otherwise poorly preserved in surface exposures. The potential for high-resolution paleoclimatic reconstruction from coring of deep-time loess strata in particular remains severely underutilized.
1 23













1 23
Your article is protected by copyright and
all rights are held exclusively by Springer-
Verlag Berlin Heidelberg. This e-offprint is
for personal use only and shall not be self-
archived in electronic repositories. If you wish
to self-archive your article, please use the
accepted manuscript version for posting on
your own website. You may further deposit
the accepted manuscript version in any
repository, provided it is only made publicly
available 12 months after official publication
or later and provided acknowledgement is
given to the original source of publication
and a link is inserted to the published article
on Springer's website. The link must be
accompanied by the following text: "The final
publication is available at link.springer.com”.
1 3
Int J Earth Sci (Geol Rundsch)
DOI 10.1007/s00531-014-1070-1
ORIGINAL PAPER
The paleoclimatic and geochronologic utility of coring red beds
and evaporites: a case study from the RKB core
(Permian, Kansas, USA)
Gerilyn S. Soreghan · Kathleen C. Benison ·
Tyler M. Foster · Jay Zambito · Michael J. Soreghan
Received: 23 March 2014 / Accepted: 7 August 2014
© Springer-Verlag Berlin Heidelberg 2014
from coring of deep-time loess strata in particular remains
severely underutilized.
Keywords Red beds · Core · Paleoclimate · Permian ·
Loess · Evaporites · Pangaea
Introduction
Drill core is well recognized as a key data set for recon-
structing climate records. Drilling marine sediments has
been long exploited to clarify climate dynamics and atmos-
pheric-oceanic linkages, especially for the Neogene, but
extending even to the Cretaceous (e.g., IODP 2008). Con-
tinental climate reconstructions also benefit greatly from
drill core data, and core from modern, long-lived lacustrine
systems in particular have facilitated major advances in
understanding climate dynamics of Earth’s recent record,
from the tropics to the poles (e.g., Lowenstein et al. 1999;
Scholz et al. 2007; Cohen 2011; Melles et al. 2012).
Despite these significant advances, the need remains for
highly resolved (orbital-forcing-scale) records that can shed
light on continental climate in Earth’s deep-time record
and reveal the full depth of the dynamic range of Earth’s
climate system (NRC 2011; Soreghan and Cohen 2013).
The orbital-scale record recovered from Triassic–Jurassic
strata of the largely lacustrine Newark basin system (Olsen
et al. 1996) exemplifies the potential for obtaining highly
resolved deep-time records from continental successions,
as does the promise of emerging results from Mesozoic
and lower Cenozoic strata of the Colorado Plateau Coring
Project (Olsen et al. 2010), and Bighorn Basin Coring Pro-
ject (Clyde et al. 2013), respectively. However, continuous
coring of continental red beds is relatively uncommon, as
these facies have little economic value, and have been long
Abstract Drill core is critical for robust and high-res-
olution reconstructions of Earth’s climate record, as well
demonstrated from both marine successions and modern
long-lived lake systems. Deep-time climate reconstructions
increasingly require core-based data, but some facies, nota-
bly red beds and evaporites, have garnered less attention
for both paleoclimatic and geochronologic analyses. Here,
we highlight studies from the Rebecca K. Bounds (RKB)
core, a nearly continuous, >1.6 km drill core extending
from the Cretaceous to the Mississippian, recovered from
the US Midcontinent by Amoco Production Company in
1988, and serendipitously made available for academic
research. Recent research conducted on this core illustrates
the potential to recover high-resolution data for geochro-
nologic and climatic reconstructions from both the fine-
grained red bed strata, which largely represent paleo-loess
deposits, and associated evaporite strata. In this case, avail-
ability of core was instrumental for (1) accessing a continu-
ous vertical section that establishes unambiguous superpo-
sition key to both magnetostratigraphic and paleoclimatic
analyses, and (2) providing pristine sample material from
friable, soluble, and/or lithofacies and mineralogical spe-
cies otherwise poorly preserved in surface exposures. The
potential for high-resolution paleoclimatic reconstruction
G. S. Soreghan (*) · T. M. Foster · M. J. Soreghan
School of Geology and Geophysics, University of Oklahoma,
100 E. Boyd St., Norman, OK 73019, USA
e-mail: lsoreg@ou.edu
K. C. Benison
Department of Geology and Geography, West Virginia University,
98 Beechurst Ave., Morgantown, WV 26506, USA
J. Zambito
Wisconsin Geological and Natural History Survey,
3817 Mineral Point Rd., Madison, WI 53705, USA
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
dismissed as of little use for paleoclimatic reconstruction,
despite limited detailed study. Many of the proxies most
commonly applied to paleoclimatic reconstructions in, e.g.,
modern/recent, organic-rich lacustrine units (e.g., organic
biomarkers, pollen, and redox-sensitive transition metals)
are considered to be compromised in oxidized systems. Yet
examples of remarkable preservation in red bed strata do
occur, such as the presence of (1) superparamagnetic mag-
netite and the accompanying magnetic susceptibility record
in Permian red beds (Soreghan et al. 1997) and (2) pristine
pollen and spores in Cenozoic and Pennsylvanian red beds
(Benison et al. 2011; Oboh-Ikuenobe and Sanchez Botero
2013; Sánchez Botero et al. 2013). Moreover, in red bed—
evaporite successions, the evaporites—if protected from
dissolution and alteration caused by post-depositional infil-
tration of dilute waters—can yield a variety of high-reso-
lution quantitative and qualitative paleoclimate data (e.g.,
Benison and Goldstein 1999; Sánchez Botero et al. 2013;
Zambito and Benison 2013). In addition, sedimentary
structures and primary fluid inclusions in Cenozoic and
Permian bedded halite can represent arid climate flooding-
evaporation-desiccation cycles and air temperature prox-
ies, respectively (Benison and Goldstein 1999; Lowenstein
et al. 1999; Zambito and Benison 2013). Paleoclimate
reconstructions from such strata, however, are dependent
upon obtaining cores drilled using methods suitable for
extracting intact evaporites and (commonly friable) fine-
grained red beds. To date, such specialized coring has been
rare, but yields remarkable data when accomplished.
The purpose of this contribution is to highlight the
importance of continuous core to develop a detailed geo-
chronologic framework and paleoclimatic record from
generally overlooked facies—in this case fine-grained red
beds and associated evaporites, even from deep-time suc-
cessions. Fortuitous access to drill core in this case ena-
bled study of (1) a stratigraphically complete section, key
for assessing a magnetic reversal stratigraphy, and (2) an
unaltered section, containing phases (evaporites, clays)
otherwise easily dissolved or compromised by traditional
drilling methods and by late-stage near-surface diagenesis.
Obtaining a continuous vertical section that establishes
unambiguous superposition is particularly critical for this
expansive, low-relief region characterized by severely lim-
ited outcrop exposures. Furthermore, the fine-grained red
beds highlighted here record an example of paleo loess, a
facies type with the potential to rival deep-sea and lami-
nated lacustrine sediments in their ability to archive high-
resolution paleoclimatic data (Liang et al. 2012), but little
exploited for Earth’s deep-time climate record. Hence, we
also highlight the vast paleoclimatic potential offered by
coring deep-time loess.
Geologic setting
Midcontinent North America during the Permian was bor-
dered by a series of orogenic systems associated with the
final assembly of Pangaea: the Ouachita-Marathon sys-
tem southward, the Appalachian system eastward, and
remnant uplifts of the Ancestral Rocky Mountains to the
west, southwest and northwest (Johnson 1978; Kluth and
Coney 1981; Slingerland and Furlong 1989; Fig. 1). This
greater region, extending from southern Saskatchewan to
Texas (north–south) and Wyoming through Kansas (west–
east), preserves a variable but commonly thick (102–103 m)
Lower-Middle Permian section (McKee and Oriel 1967;
Fig. 1 Paleogeographic map
of North America for Middle
Permian (265–255 Ma) time,
from Blakey (Colorado Plateau
Geosystems, http://cpgeosys
tems.com), modified to show
orogenic belts of the Appala-
chian, Ouachita-Marathon, and
ARM (Ancestral Rocky Moun-
tains) systems. Permian red bed
strata are preserved across a
broad region of Midcontinent
North America (see Fig. 2).
Note paleoequator line and
paleoequator
Region
of
Midcontinent
Permian
Red Beds
500 km
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
Walker 1967) that extends across various basins and posi-
tive areas and reflects as-yet poorly understood subsidence
that is perhaps related to post-orogenic and/or far-field
effects (Soreghan et al. 2012).
Lithofacies in the Midcontinent Permian range from red
beds interbedded with marine limestone low in the section,
to entirely continental red beds and evaporites high in the
section (McKee and Oriel 1967). This transition reflects a
well-documented eustatic and climatic shift driven by (1)
the evolution from the Permo-Carboniferous icehouse cli-
mate to full greenhouse conditions in the Permo-Triassic
(Frakes 1979), with attendant high-frequency glacioeustasy
detectable predominately low in the section (e.g., Heckel
2008) and (2) the gradual emergence of the Pangaean
supercontinent, as relative sea level reached its Phanero-
zoic minimum near the end of the Permian (Ross and Ross
1988, 1994, 1995). The latter trend resulted in the predomi-
nance of continental over marine deposition through most
of the Permian in the Midcontinent and indeed globally
(e.g., Golonka and Ford 2000).
The Midcontinent USA was situated in the western
equatorial region (~5–15°N; e.g., Golonka et al. 1994; Sco-
tese 1999; Kent and Muttoni 2003; Loope et al. 2003) of
Pangaea throughout Permian time, yet the climate was gen-
erally arid, and inferred to have become increasingly arid
from Pennsylvanian through Permian time (e.g., Parrish
1993), although the forcing for this trend remains debated
(Tabor and Poulsen 2008, and references therein). Both
models and data indicate the onset of monsoonal circula-
tion across Pangaea by early Permian time (e.g., Robinson
1973; Parrish and Peterson 1988; Kutzbach and Gallimore
1989; Parrish 1993; Soreghan et al. 2002; Tabor and Mon-
tanez 2002).
Within the greater study region of Kansas and Okla-
homa, the Permian interval includes, from base to top, the
Council Grove, Chase, Sumner, and Nippewalla, groups
(and three post-Nippewalla Formations in Kansas that are
likely equivalent to the Quartermaster Group in Oklahoma;
Fig. 2). Zambito et al. (2012) presented the detailed stra-
tigraphy of the studied interval and difficulties regarding
correlation in the region. Regionally, the Council Grove
and Chase groups consist generally of lithologically mixed
(marine) carbonate and (continental) siliciclastic “cyclo-
thems”, whereas the overlying groups consist predomi-
nantly of red siliciclastic strata and bedded and displacive
evaporite strata. The loss of a marine signal up section in
this stratigraphy, together with laterally discontinuous out-
crops that are poorly preserved owing to late-stage meteoric
dissolution has long stymied efforts of stratigraphic dating
and correlation. Compounding this, few complete (and well
preserved) cores exist, owing to the limited economic inter-
est in this shallow section. Hence, detailed study of envi-
ronments and paleoclimate of the Midcontinent Permian,
and—by extension—the record of greater western equato-
rial Pangaea—remains severely hampered.
Background and methods
The Rebecca K. Bounds No. 1 (RKB) core was drilled
by Amoco Production Company in 1988, in westernmost
N500 km
Permian red beds Permian red beds
and evaporites
A
B
Permian
Lopingian
Changhsingian
Wuchiapingian
Capitanian
Wordian
Roadian Nippewalla
Sumner
Chase
Council Grove
Big Basin
Day Creek &
Stage Kansas
Lithostratigraphy
Age (Ma)SeriesPeriod
Kungurian
Artinskian
Sakmarian
Asselian
Guadalupian
Cisuralian
295.0
290.1
283.5
272.3
268.8
265.1
259.8
254.1
Whitehorse Fms
Group
Group
Group
Group
(part)
Core locality
252.2
298.9
Fig. 2 a Simplified map showing areal extent of Permian red beds
of the Midcontinent US (modified from Benison and Goldstein 2001;
originally from Walker 1967) and b chronostratigraphy for the RKB
core. Age designations for the Sumner and older groups are from
Sawin et al. (2008) and West et al. (2010; originally Norton 1939),
converted to international timescale designations using the most
recent Permian timescale (Gradstein et al. 2012 for the conversion
of stages; and Shen et al. 2013 for the latest dates). Chronostratig-
raphy for the Sumner Group and above reflect new magnetostrati-
graphic age assignments (Foster et al. 2014) discussed in the text, and
detailed in Figs. 4 and 5
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
Kansas, USA (Figs. 2, 3). The core was drilled as an experi-
mental project, primarily to test the capabilities of Amoco’s
(then) newly developed “SHADS” (Slim-Hole Advanced
Drilling System) rig (Walker and Millhein 1989), and sec-
ondarily to buttress biostratigraphic work on composite
standards for the Paleozoic and Mesozoic. The SHADS
technology included not only the capability of drilling and
coring rapidly and continuously, but significant on-site core
analysis using a modular, portable facility. The drilling
of the RKB core required 38 days and US$ 591,000, and
resulted in recovery of 1,656 m of continuous core extend-
ing from the Cretaceous nearly through the Mississippian,
with >90 % recovery overall (Dean and Arthur 1998; Dean
et al. 1995; Wahlman pers. commun. 2014). Amoco sub-
sequently analyzed the core primarily for Mississippian-
Pennsylvanian foraminiferal and fusulinid biostratigraphy
(Wahlman and Groves pers. commun. 2014). According
to Amoco internal reports on the RKB core, the young-
est biostratigraphically significant marine fossils from the
Paleozoic interval were latest Pennsylvanian (Gzhelian)
fusulinids (Wahlman pers. commun. 2014). The typically
cyclic upper Pennsylvanian section exhibits a gradual loss
of normal marine deposition, with much of the upper-
most Pennsylvanian composed of dolomitic to anhydritic
restricted-marine facies. Buijs and Goldstein (2012) and
Dubois et al. (2012) conducted petrographic observations
Fig. 3 Location and summa-
rized stratigraphic column for
the Amoco Rebecca K. Bounds
No. 1 core, Greeley County,
Kansas. Red star on inset map
in upper left shows approximate
core location. See Zambito
et al. (2012) for a more detailed
lithological log
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
and well log analyses on the Carboniferous-Lower Permian
section. Several publications have focused on the Mesozoic
strata of the RKB core (e.g., Arthur 1993; Dean et al. 1995;
Dean and Arthur 1998). As oil prices continued to fall
through the late 1980s and 1990s, the SHADS technology
was ultimately abandoned (Wahlman pers. commun. 2014).
Splits of the core are now housed at the Kansas Geological
Survey core repository in Lawrence, Kansas, as well as the
US Geological Survey in Denver, Colorado.
The Permian section of the core consists predominantly
of fine-grained red beds with varying amounts of intergran-
ular and displacive halite cement, as well as bedded evapo-
rite strata (Fig. 3). Core recovery in this interval (~488 to
~1,034 m subsurface) is 99.1 %, reflecting the effective
use of a drilling fluid engineered for water-sensitive facies
(Zambito et al. 2012; Benison et al. 2013). Despite the
excellent recovery, this interval remained virtually unstud-
ied until very recently (Benison et al. 2013; Zambito et al.
2012; Foster 2013; Kane 2013, Zambito and Benison 2013;
Foster et al. 2014), as the facies are barren of marine fossils
typically used for biostratigraphic zonation in this inter-
val, and the fine-grained red bed and evaporite facies were
ignored for other detailed analyses.
Beginning in 2011, the section of the core from ~1,034
to 488 m (Middle Permian) was measured, logged, and
sampled to establish lithostratigraphy through the tar-
geted intervals (~Lower Permian–Middle Permian; see
Zambito et al. 2012 for detailed stratigraphy). In addition,
description and sampling were conducted at cm-scale res-
olution through several intervals chosen to focus on (1)
fine-grained red beds of selected units and (2) pristine
evaporite strata. For magnetostratigraphic analysis, sam-
ples marked with their stratigraphic “up” direction were
collected at ~75 cm intervals (avoiding apparent bio- or
pedoturbated samples) through the upper 105 m of the
Permian interval and subjected to thermal demagnetiza-
tion. Inclination data were used to track changes in incli-
nation to assess magnetic reversals (see Foster 2013 for
details of the magnetostratigraphic analysis). Two hun-
dred and thirty thin sections were made using vacuum
impregnation of epoxy. Care was taken not to heat sam-
ples or use water during thin section preparation. Thin
sections were examined with transmitted, reflected, polar-
ized, and UV–Vis light at magnification up to 2000x. Hal-
ite was prepared for fluid inclusion analyses by cleaving
to mm-scale chips with a razor blade. Fluid inclusion
heating runs were conducted on a USGS-modified gas-
flow fluid inclusion stage (Benison and Goldstein 1999)
and a Linkam THMSG600 fluid inclusion stage (Zambito
and Benison 2013). Twenty-six additional samples rep-
resentative of facies throughout the study interval were
analyzed by X-ray diffraction for general mineral iden-
tification. Continuous core scans (e.g., XRF) were not
performed as no facilities nor funding were available for
such work.
Data acquisition enabled by coring
Data collected as part of the research on the Permian of
the RKB core include, to date, petrography, sediment geo-
chemistry, detrital zircon geochronology, magnetostratigra-
phy, isotope geochemistry, fluid inclusion microthermom-
etry, and preliminary clay mineralogy (e.g., Benison et al.
2013; Zambito and Benison 2013; Foster 2013; Kane 2013;
Foster et al. 2014). Our goal here is to highlight selected
data collections enabled uniquely by drill core acquisition,
together with their chronologic or paleoclimatic utility, and
data acquisitions key to paleoclimatic reconstructions in
paleo-loess successions that could be done in the future on
this or analogous systems in optimally located sites. From
the RKB core specifically, we highlight results from mag-
netostratigraphy and fluid inclusion microthermometry.
Critically, none of the results could have been put into an
unambiguous stratigraphic succession without the aid of
such a high-recovery, well-preserved drill core.
Magnetostratigraphy
Magnetostratigraphy of the Permian part of the RKB core
reveals a robust reversal stratigraphy (Fig. 4a). The Zijder-
veld diagrams (Fig. 4b) show the presence of two magnetic
components: a low-temperature component (0–250 °C) rep-
resenting a modern viscous remanent magnetization (VRM),
and a higher-temperature (typically ~550–675 °C, or ~450–
550 °C) component. Based on abundant unblocking tem-
peratures above 580 °C, we infer the DRM/CRM (detrital
remanent magnetization, chemical remanent magnetization)
resides in hematite. Analysis of these data reveals a robust
reversal stratigraphy; the presence of this sequence of rever-
sals suggests the magnetization is either a DRM or an early
CRM. That is, in either case, the magnetization had to be
early in order to record the reversal stratigraphy. Addition-
ally, the magnetic inclination data are consistently low (Fos-
ter 2013; Foster et al. 2014), and thus most consistent with
early acquisition of the magnetization. The oldest reversal
occurs near the contact between the Dog Creek Shale and
the Whitehorse Formation at a depth of ~560 m, where the
inclination changes from reversed to normal (Fig. 4). Alto-
gether, three reversed polarity events (from bottom to top;
event 1: avg. inclination = 15.8°, SD = 9.3°; event 2:
avg. inclination = 24.4°, SD = 17.8°; and event 3: avg.
inclination = 7.9°, SD = 9.3°) and two normal polarity
events (event 1: avg. inclination = 16.5°, SD = 8.9°; and
event 2: avg. inclination = 12.3°, SD = 7.6°) occur through
the ~105 m of core leading up to the end Permian, clearly
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
placing the Whitehorse and Big Basin Formation in post-
Kiaman time (Figs. 4, 5). In addition, an average inclination
of 16.3° throughout the sampled interval is consistent with
inclination values for the middle to late Permian seen in the
variation of inclination with time for the study site (Fig. 6),
providing further support for a post-Kiaman age. Combin-
ing these data with previous work on chronostratigraphy of
the section (Denison et al. 1998; Foster et al. 2014) substan-
tially refines the previously proposed chronostratigraphic
placements for the midcontinent region (Fig. 5). In addition
to revising the timing and duration of deposition, paleo-
magnetic data were used to identify paleolatitude estimates.
Consistently, low inclination values throughout the sampled
interval indicate an average paleolatitude of 6–10°N (Fig. 6;
van der Voo (1993), confirming the lower end of previously
cited ranges.
Evaporite paleothermometry
Cores of Permian red beds and evaporites throughout the
US Midcontinent from a depth window of ~300–2,100 m
are well preserved. In contrast, outcrops and deposits
within ~300 m of the surface have undergone late-stage
dissolution and alteration from recent groundwaters (Beni-
son and Zambito 2013) such that detailed petrography can
only be accomplished well with the aid of drill core. The
recognition of unaltered bedded halite, displacive halite,
and halite cements in red beds (Fig. 7) yields complete and
well-preserved core necessary for assessing high-resolution
depositional and early diagenetic conditions. For example,
the displacive halite lithology (aka “chaotic halite” in some
older literature), composed of red mudstone with randomly
oriented large halite crystals, forms syndepositionally in
groundwater-saturated saline mudflats adjacent to ephem-
eral saline lakes in arid climates (Benison and Goldstein
2001; Benison et al. 2007; Lowenstein and Hardie 1985).
This lithology is the most abundant lithology in the Flow-
erpot Shale in the RKB core (Benison et al. 2013). How-
ever, in outcrop, the displacive halite crystals have been
dissolved during late-stage, near-surface diagenesis by low-
salinity groundwaters. Therefore, this saline mudflat litho-
facies appears as massive mudstone, with only rare halite
hopper crystal casts, molds, and pseudomorphs as evidence
of the original depositional environment.
Primary fluid inclusions in bedded halite and displacive
halite from the RKB core are well-preserved remnants of
Permian surface waters and groundwaters. They can be
tested with microthermometric methods and various geo-
chemical analyses to yield temperatures, water salinities,
and even water pH (Benison 2013). Homogenization of
artificially nucleated vapor bubbles in primary fluid inclu-
sions in chevron halite measure the temperature of Permian
shallow (less than ~0.5 m) saline water at the time that the
halite was growing. Because shallow surface waters have
approximately the same temperature as local air tempera-
ture, these homogenization temperatures can be considered
proxies for ancient air temperatures. Primary fluid inclu-
sions in chevron halite define daily growth bands (Roberts
and Spencer 1995; Benison and Goldstein 1999). Careful
petrography enables high-resolution stratigraphic control of
air temperature proxies at daily scales. This yields diurnal
temperature ranges over days to weeks.
Zambito and Benison (2013) measured homogeniza-
tion temperatures from primary fluid inclusions from 15
beds of chevron halite from the undifferentiated Salt Plan
Formation/Harper Sandstone, the Cedar Hills Sandstone,
and the Blaine Formation in the RKB core. Temperatures
ranged from 7 to 73 °C. The maximum diurnal temperature
range was 32° C. Trends in homogenization temperatures
from base to near the top of the Nippewalla Group showed
Stratigraphy
(RKB core)
Whitehorse Formation
Dog Creek
Shale
Big Basin
Formation
500
510
520
530
540
550
560
570
580
Depth (m)
Inclination ( )
o
0-50 50
Polarity
< 267 Ma
N, Up
W
N, Up
E
#1734
#1836
VRM
NRM NRM
VRM
A
B
650 C
o
0 C
o
550 C
o
650 C
o
0 C
o
200 C
o
Fig. 4 a Stratigraphy of middle Permian red bed strata of the RKB
core sampled for magnetostratigraphy, showing resultant magnetic
inclination data, and magnetic polarity data of middle Permian red
beds from western Kansas. Black represents normal polarity and
white represents reversed polarity. b Typical Zijderveld demagnetiza-
tion plots of normal (left) and reversed (right) polarity. Present-field
magnetization, VRM viscous remanent magnetization, NRM natural
remanent magnetization
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
warming and then cooling. This is quantitative, high-res-
olution paleoweather and paleoclimate data that strongly
suggests extremely warm temperatures in western Pangea
during the mid Permian.
Discussion: Why core continental red beds
and evaporites?
The problem of time
Midcontinent Permian red beds are predominantly fine-
grained, and poorly lithified, owing to generally shallow
burial (Carter et al. 1998; Hemmerich and Kelley 2000;
Foster et al. 2014). Moreover, the low relief of the region
in combination with the commonly low stratigraphic dips
(~0.5°) has stymied outcrop-based studies of these units.
Widely dispersed outcrops throughout the region expose
only a few meters to a few tens of meters of section, and
tend to be highly weathered and unstable owing to the fri-
able, fractured, fine-grained, and evaporitic character of the
facies (Fig. 8). Moreover, a fundamental obstacle that has
prevented detailed study of the red bed-dominated section
of the RKB core, and indeed outcrop systems throughout
the North American midcontinent, is lack of a reasonable
age model, owing to the paucity of biostratigraphically sig-
nificant fauna. The resulting dearth of temporal resolution
has impeded progress in regional and global correlations,
and thus integration of the vast amounts of data preserved
in these strata.
Magnetostratigraphy, although of limited use for the
Early Permian, provides a potentially powerful dating
PERMIAN
Cisuralian Guadalupian
Kungurian Roadian Wordian Capitanian
System
Series
Stage
Geomagnetic
Polarity Primary
Kiaman Reversed-polarity Superchron
South Central
Kansas
North - Central
Oklahoma
Group Formation
Nippewalla
Sumner
Harper SS.
Salt Plain
Cedar Hills SS.
Flowerpot Sh.
Blaine
Whitehorse
Day Creek Dol.
Big Basin
Hennessey
El Reno Whitehorse
Duncan SS.
Flowerpot
Blaine
Marlow
Rush Springs
SS.
Cloud Chief
Norton (1939) Johnson et al. (1989)
Group Formation
Dog Creek
Dog Creek Sh.
Group Formation
Nippewalla
Harper SS.
Salt Plain
Cedar Hills SS.
Flowerpot Sh.
Blaine
Whitehorse
Day Creek Dol.
Big Basin
Dog Creek
Proposed
Chronostratigraphy
(this study)
Fig. 5 a Chronostratigraphy and (new) magnetostratigraphy of mid-
dle Permian red beds from western Kansas. Black represents normal
polarity and white represents reversed polarity. South Central Kan-
sas column is from Norton (1939; see also West et al. 2010), Swin-
eford (1955), Ham (1960), and Baars (1990). For comparison, we
also show north central Oklahoma stratigraphy (Johnson 1989a, b).
Shaded area represents ~105 m of red beds sampled for palaeomag-
netic data. Note that the Kansas Geological Survey places this inter-
val entirely within the Kiaman Superchron (i.e., >267 Ma). Shaded
bars displays the positions of normal polarity events found in the
RKB core
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
and correlation tool for the Middle-Late Permian, as the
first reversal after the Kiaman Superchron occurred in the
Wordian (~267 Ma; Steiner 2006; stratigraphy.org; Fig. 5).
Building a magnetostratigraphic framework, however,
requires acquiring samples that are (1) sufficiently indu-
rated and free of surface weathering to enable effective
sampling and thermal demagnetization and (2) collected
in a long, continuous section with unambiguous superposi-
tion to enable construction of a robust reversal stratigraphy.
This cannot be done in the Midcontinent Permian without
core, owing to the limits of the surface exposures.
Detrital zircons of volcanic origin, however, can be
used to constrain the depositional ages of sedimentary
units that are otherwise poorly dated (e.g., Dickinson and
Gehrels 2009; Soreghan et al. 2008, 2014) driven in part
by advances in geochronologic methodology. For exam-
ple, the ability to screen the U–Pb ages of a large number
of grains through laser-ablation methods, followed by ID-
TIMS (Isotope-Dilution Thermal Ionization Mass Spec-
trometry) analysis of the youngest grains in the sampled
population can provide high-precision ages of grains that
may correspond to the depositional age of the deposit. The
derived age of the grain represents a maximum age of dep-
osition; i.e., the grain must be as old as the sampled hori-
zon, although the sampled horizon could be younger. This
has been done on outcrop studies, but this method can be
used with additional benefit in studies of continuous core as
it allows the determination of maximum age of deposition
at various horizons, eliminating the need to sample forma-
tions from spatially disparate outcrops where relative age
information is not known a priori. The only potential draw-
back to this method is that the volume of sample needed
from the core can be substantial; however, in our work
(Kane 2013), a core split (half of 8.5 cm diameter core) of
~50 cm length of sandstone, siltstone, or mudstone yielded
a sufficient number of zircons, and still enabled preserva-
tion of the archival half, as well as discrete sampling for
auxiliary (e.g., thin section) analyses.
Red is not dead: the paleoclimatic value of red beds—
especially paleo-loess
Lake systems have long been considered an ideal envi-
ronment to tap for paleoclimatic reconstructions, as (per-
manent) lakes archive a continuous or near-continuous
record that enables analysis at high temporal resolution of
multiple metrics of paleoclimate, including proxies recon-
structed from lithology, magnetism, geochemical and iso-
topic signals (e.g., Brigham-Grette et al. 2007; Cohen
2011). In recognition of this, research drilling for conti-
nental paleoclimatic reconstructions has focused in many
-80
-60
-40
-20
0
20
40
60
80
050100150200250300350400450500
Inclination ( °)
Age (Ma)
Perm
Carb
Dev
Si
Ord Tr Ju Cret P
Fig. 6 Plot of the expected inclination through time for the north-
ern study site. The gray lines represent error. Data from van der Voo
(1993)
Fig. 7 The three main types of halite in the Rebecca K. Bounds core.
a Bedded halite with abundant red mud, (714.4 m; 2,344; thick sec-
tion, transmitted light), b Displacive halite (773.3 m; 2,537; thin sec-
tion, transmitted light). c Intergranular halite cement in sandstone
(738.2 m; 2,422.25; thin section, transmitted light)
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
cases on coring of modern, long-lived lake systems (e.g.,
Lake Malawi, Lake El’gygytgyn, Lake Titicaca, Lake Peten
Itza) with records extending in some cases to the Pliocene
and even late Miocene (Cohen 2011). The time continuum
of the deposits is particularly useful, but lake sediments
also offer the potential to utilize several different types of
organic carbon-based proxy analyses, such as compound-
specific carbon isotopic ratios, and tetraether-based proxies
(e.g., TEX-86 and isoprenoid glycerol dialkyl tetra ether;
e.g., Tierney 2010).
Red beds occur in both marine and continental set-
tings, but are far more common in the continental record.
The red color primarily reflects (oxidized) hematite con-
tent, meaning little to no organic carbon may remain, and
thus minimal potential for measurement of climate proxies
based on such material. Some studies on red beds associ-
ated their presence with particular paleoclimatic conditions
(review in Dubiel and Smoot 1995; Hu et al. 2014), such
as warm arid settings, but other work has demonstrated
that red beds form in a variety of climatic settings, from
tropical to desert, suggesting caution in paleoclimatic inter-
pretations of a red color (Dubiel and Smoot 1995; Sheldon
2005). Dubiel and Smoot (1995) noted that red bed for-
mation reflects several conditions, including the presence
of (1) small amounts of precursor organic matter (Myrow
1990), (2) abundant labile material such as mafic minerals
and lithic fragments (e.g., Walker 1967, 1976), and (3) oxi-
dizing conditions (Walker 1967; Turner 1980) Although the
red color is not necessarily indicative of paleoclimate, vari-
ous types of sedimentologic, geochemical, magnetic, and
paleontologic criteria support detailed paleoclimatic recon-
struction from red beds (details below).
Particularly critical for paleoclimatic reconstruction is
preservation of a continuous record of surface paleoenvi-
ronments. Although challenging for some types of con-
tinental red beds (e.g., fluvial), paleo-loess systems are
particularly well suited in this regard. The Chinese Loess
Plateau (CLP) is considered an excellent archive of conti-
nental paleoclimate—directly comparable in resolution to
ice-core and deep-marine archives (e.g., Liu 1985; Kukla
and An 1989; Bloemendal et al. 1995; Liu et al. 1999;
Ding et al. 2002). In many settings, loess accumulates very
quickly, producing high temporal resolution (e.g., 1 m/ky
in Rhine Valley) (Hatte et al. 2001; Lang et al. 2003) that
rivals or exceeds any other continental depositional system,
and responds directly to atmospheric conditions.
Study of loess as a high-resolution paleoclimate archive
has long been conducted for the Quaternary record, but
loess remains an under-utilized archive for Earth’s deep-
time record, despite increasing recognition of deep-time
loess deposits (e.g., Johnson 1989a, b; Soreghan 1992;
Evans and Reed 2007; Soreghan et al. 2008). The Late Car-
boniferous-Permian record appears to be particularly rich
in occurrence and preservation of thick and widespread
loess deposits (Soreghan et al. 2008), and many of the fine-
grained red beds of this age in the North America midcon-
tinent and elsewhere have been recently reinterpreted as
paleo-loess deposits (e.g., Sweet et al. 2013; Dubois et al.
Fig. 8 Photographs of the studied Permian section in surface expo-
sures, illustrating the character of outcrop. a Permian Dog Creek
Shale exposed in central Oklahoma. b Permian Flowerpot Shale
exposed in central Oklahoma. c Permian Flowerpot Shale exhibiting
outcrop dissolution of evaporites
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
2012; Giles et al. 2013), including units of the RKB core
(Foster 2013; Foster et al. 2014). In addition to loess (and
associated paleosol) deposits, these facies include lake
deposits as well—but shallow and ephemeral saline lakes,
rather than deep, oxygen-poor systems.
Recognition of the widespread occurrence of loess and
(saline) lake deposits over a broad region of the midconti-
nent promises the potential of very high-resolution climatic
reconstruction, given the possibility of continuous sampling
enabled by coring. Loess and associated deposits (e.g.,
paleosols, saline lake deposits) house enormous poten-
tial for paleoclimatic reconstruction. Analogizing again to
the CLP, various attributes of loess (e.g., grain size, mag-
netic susceptibility, and geochemistry) have been mined to
reconstruct climatic parameters that include atmospheric
circulation (wind velocity and direction), seasonality, and
precipitation (e.g., Zhou et al. 1990; An et al. 1991; Bloe-
mendal et al. 1995; Liu et al. 1995; Ding et al. 2002; Sun
2002; Balsam et al. 2004; Vandenberghe et al. 2004; Hao
and Guo 2005; Chen et al. 2006), at resolutions extending
to millennial. Many of these same metrics are preservable
and measurable in deep-time loess deposits and have ena-
bled climate reconstructions ranging to sub-precessional
scales (Soreghan et al. 2014).
As an example, the Maroon Formation, a Permian
loessite-paleosol succession in central Colorado is well
exposed along road cuts and has been extensively sam-
pled (Johnson 1989a, b; Soreghan et al. 1997; Tramp et al.
2004; Soreghan et al. 2014). Tramp et al. (2004) noted that
the alternating loessite-paleosol couplets exhibit similari-
ties to the CLP both in lithologic character and in appar-
ent temporal patterns of magnetic susceptibility. Sub-meter
sampling of the 700 m section showed that magnetic sus-
ceptibility values are higher in deeply red-colored, finer-
grained beds, whereas values are lower in orange, silty
units. These changes are interpreted to reflect alternat-
ing wet-dry phases linked to climate swings within the
late Paleozoic icehouse. To further explore these patterns,
Soreghan et al. (2014) sampled a 12 m interval of this same
section of the Maroon Formation on a 10-cm scale for geo-
chemical, grain size, and magnetic susceptibility trends.
They documented a robust negative correlation (r2 = 0.9)
between grain size (inferred from image analysis of quartz
grains; Soreghan and Francus 2004), and magnetic sus-
ceptibility values (Fig. 9). The variations in quartz grain
size are interpreted to record changes in wind intensity
(and/or source proximity) with finer, more iron-rich sedi-
ment deposited during times of reduced winds (or from
further distances) and coarser, more quartzose (less iron-
rich) sediment deposited during times of stronger, seasonal
winds (Soreghan et al. 2014). The finer-grained beds show
evidence of pedogenesis and are interpreted to represent
wetter conditions that grade downward into the coarser,
orange units inferred to be loessite deposited during more
arid times. However, the nature of the transitions and the
internal stacking of these loessite-paleosol couplets suggest
that they represent high-frequency fluctuations in wind pat-
terns and wind intensity. The thicker loessite units, capped
by thicker, well-developed paleosols show an abrupt fining
in grain size with the coarsest sediment at the transition;
thinner loessite-paleosol couplets show a more gradual fin-
ing and less variation overall. These alternations, and their
nested variability, typical of the entire exposed 700 m of
the Maroon Formation may reflect sub-Milankovitch vari-
ability. However, attempting to create a continuous record
using surface exposures is untenable at the resolution nec-
essary to delineate this variability. A continuous core, bol-
stered by further refinement of the grain size to magnetic
susceptibility correlation, would facilitate reconstruction
of a very high-resolution record of wind regimes and thus
atmospheric circulation from deep time.
Furthermore, continuous acquisition of metrics such as
magnetic susceptibility, XRF-based geochemistry, spectral
reflectance, and other data readily acquired from core using
rapid core scanning and/or high-resolution point sampling
are ideal for quantitative assessments of cyclostratigraphy,
which further improves age models and potentially enables
resolution of climatic evolution down to the 10 ky scale
(e.g., Olsen et al. 1996; Sur et al. 2010).
Paleoclimatic information can also be determined from
rocks lacking typical easily interpretable sedimentological
features based on changes in clay mineralogy. Such pale-
oclimatic interpretations from clay mineralogical signa-
tures must first be assessed for influence of non-climatic
influences such as extreme water chemistry, as well as
changes in source area, tectonic forcing, and sedimentolog-
ical sorting processes on detrital clays, and diagenetic pro-
cesses. The climatic signal can potentially be interpreted
from very subtle changes in the ratios of clays represent-
ing physical weathering (illite/chlorite), seasonal chemical
weathering (smectite), and persistent wet conditions (kao-
linite) (Arostegi et al. 2011). Clay mineralogy from drill
core ensures that surface weathering has not degraded the
signal.
Climate and weather revealed in evaporites
Well-preserved ancient evaporites, accessible only in cores,
provide paleoclimate data in two distinct ways. Petro-
graphic documentation of halite and gypsum crystal types
and sedimentary features lead to informed interpretations
of depositional environments. Because most evaporite
depositional environments are sensitive to climate, pet-
rographic observations provide qualitative information
about paleoclimate, such as relative aridity (i.e., Benison
et al. 2007; Lowenstein and Hardie 1985). Secondly, fluid
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
inclusion data from bedded halite yield high-resolution,
quantitative records of paleoclimate. Homogenization of
artificially nucleated primary fluid inclusions are prox-
ies for past air temperatures (Roberts and Spencer 1995).
Homogenization temperatures measured from base to
top of individual growth bands in chevron halite allows
for interpretation of minimum daily temperature ranges,
whereas homogenization temperatures from successive
chevron growth bands suggest diurnal temperature ranges
(Benison and Goldstein 1999). Longer-scale trends in air
temperatures can be resolved by comparison of homog-
enization temperatures among individual beds of hal-
ite. In addition, chemical compositions of primary fluid
inclusions in ancient halite can document past surface
Fig. 9 a Detailed measured
section of a 12 m interval of
the Permian Maroon Forma-
tion near Basalt, Colorado
(see Soreghan et al. 2014
for location). Bulk magnetic
susceptibility values consist-
ently increase at the deeper red
(rb= “red-brown”) horizons
bearing lithologic indicators
of pedogenesis (blocky peds,
root traces) and decrease within
the orange-colored (org=
“orange”) non-pedogenically
altered loessite horizons. b
Apparent grain area based on
image analysis of ~800 quartz
grains per sample normalized
to their stratigraphic position
relative to inferred paleosol tops
within four loessite-paleosol
couplets targeted within the
measured section. The samples
from the thickest couplets
exhibit a marked increase in
apparent grain size just below
the paleosol top, then fine
abruptly at the paleosol horizon
(dashed line with arrow),
whereas samples from thinner
couplets exhibit a more gradual
decrease from loessite to
paleosol (solid line with arrow).
These appear to reflect changes
in the variability of monsoon
circulation on sub-Milankovitch
scales (see Soreghan et al.
2014). Modified from Soreghan,
et al. 2014
0400 800 1200 1600
.2
.4
.6
.8
1.0
Mean apparent grain
area (sq. microm.)
normalized stratigraphic position
top of paleosol
org
rb
dk
rb
dk
rb
rb
dk
rb
rb
dk
rb
rb
dk
rb
org
rb
dk
rb
1
2
3
4
5
6
7
8
9
10
11
12
rb
org
rb
dk
rb
Color
Thickness (m)
org
dk
rb
org
org
org
org
rb
org
thin couplets thin couplets
thick couplet thick couplets
root traces
blocky peds
key to symbols
discontinuous
laminations
024681012
Magnetic
Susceptibility
+-8 m3/kg)
BA
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
and groundwater chemistry. This chemical data can pro-
vide information about weathering processes that may
be dependent upon climate. Furthermore, sampling from
continuous core confers unambiguous superposition, such
that these valuable paleoclimatic data can be placed into a
proper temporal context.
Caveats and future opportunities
The mere existence of a core like the RKB offers abun-
dant opportunities, but much more could be done given
core scanning. Amoco acquired standard oil industry
(well-bore) logging during drilling, but no additional scans
were ever conducted on the core. For the purposes of high-
resolution environmental reconstruction, multi-sensing
core logging techniques (MST) offers a relatively low-
cost means to rapidly and nondestructively characterize
core samples. Such data include neutral gamma radiation,
gamma density, P-wave velocity, magnetic susceptibil-
ity, electrical resistivity, and imaging. Data acquired from
MST can significantly improve the ability to characterize
changes in lithology that can then direct more detailed and
time-consuming analyses to follow. Density and P-wave
velocity measurements can be used to generate a syn-
thetic seismogram to enable correlation of the core to seis-
mic records. MST also enable continuous digital imaging
(color spectrophotometry) of the core for archival as well
as image analyses.
Finally, although access to core provides unrivaled
opportunities to sample fresh material in unambiguous
superposition, the one-dimensionality of a core is clearly
limiting relative to the multi-dimensionality of expansive
outcrops. Where permitted by the geologic setting, there-
fore, a coring program that combines drilling with outcrop
studies remains the ideal approach. This is particularly
true for characterizing linked environmental and paleobio-
logical change, where outcrop studies may provide more
access to fossil material, and core provides the context
of continual environmental change. The challenge then
remains to correlate these records at high resolution, using
geochronological and magnetostratigraphic approaches.
An excellent example is the current attempt to catalog
the possible environmental drivers to human evolution in
East Africa, where the fossil and artifact record has long
been pursued in isolated outcrops. Tying this record to
an archive of continuous environmental change is now
occurring through a large-scale coring program (Cohen
et al. 2009). Several of the drilling sites for this project
were chosen in direct proximity to well-studied outcrops
to facilitate direct correlation of the (outcrop-based) fossil
and artifact record to the (core-based) paleoenvironmental
record.
Conclusions
The RKB core, drilled with the primary intent of testing a
drilling methodology, serendipitously aided preliminary
chronostratigraphic, and paleoclimatic refinements for the
Permian of the North American Midcontinent. The studied
units consist entirely of red bed and evaporitic strata, com-
monly dismissed as being of little utility for chronostrati-
graphic and paleoclimatic work. Yet fine-grained red beds
are ideal for magnetostratigraphic analysis and house the
potential to yield a wealth of high-resolution data for pale-
oclimatic reconstructions, especially if these strata record
paleo-loess deposition. Furthermore, contained evaporite
strata, pristine in core, can provide quantitative paleocli-
matic and even paleo-weather data. Such paleo-loess and
evaporite deposits characterize much of the Permian record
in many regions globally as well as many regions plagued
by low relief and poor outcrop exposure. In these succes-
sions, drill core is essential for (1) accessing a continuous
vertical section that establishes unambiguous superposi-
tion key to both magnetostratigraphic and paleoclimatic
analyses and (2) providing pristine sample material from
friable, soluble, and/or lithofacies and mineralogical spe-
cies otherwise poorly preserved in surface exposures. Our
work on the RKB core illustrates a fraction of the potential
that coring in such units can offer. The ability to drill in key
regions representing the most continuous sections and con-
duct continuous core scanning and auxiliary types of proxy
analyses would shed abundant light on a critical icehouse–
greenhouse transition in Earth history.
Acknowledgments We owe deep thanks to G. Kullman (Oklahoma
Geological Survey core repository) for originally suggesting the RKB
core for our study, and J. Groves, R. Scott, and especially G. Wahl-
man (all formerly of Amoco Production Company) for sharing find-
ings and history from the RKB core. We thank R. Buchanan, D. Laf-
len, and L. Watney (Kansas Geological Survey) for providing access
to and sampling of the RKB core. Funding for this research was pro-
vided by grants from the National Science Foundation (EAR-1053018
to MJS and GSS, and EAR-1053025 to KCB). We thank reviewer C.
Heil, an anonymous reviewer, and editor C. Dullo for constructive
comments on an earlier version of this manuscript.
References
An ZS, Kukla GJ, Porter SC, Xiao JL (1991) Magnetic susceptibility
evidence of monsoon variation on the Loess Plateau of central
China during the last 130,000 years. Quat Res 36:29–36
Arostegi J, Baceta JI, Pujalte V, Carracedo M (2011) Late Creta-
ceous–Palaeocene mid-latitude climates: inferences from clay
mineralogy of continental-coastal sequences (TrempGraus area,
southern Pyrenees, N Spain). Clay Miner 46:105–126
Arthur MA (1993) Cretaceous shallow drilling, US western interior:
core research, project number DE FG02-92ER14251. Techni-
cal progress report, Department of Energy, DOE/ER/14251–1,
31 pp
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
Baars DL (1990) Permian chronostratigraphy in Kansas. Geology
18:687–690
Balsam W, Ji JF, Chen J (2004) Climatic interpretation of the Luo-
chuan and Lingtai loess sections, China, based on changing iron
oxide mineralogy and magnetic susceptibility. Earth Planet Sci
Lett 223:335–348
Benison KC (2013) Acid saline fluid inclusions: examples from mod-
ern and Permian extreme lake systems. Geofluids 13:579–593.
doi:10.1111/gfl.12053
Benison KC, Goldstein RH (1999) Permian paleoclimate data
from fluid inclusions in halite. Chem Geol 154:113–132.
doi:10.1016/S0009-2541(98)00127-2
Benison KC, Goldstein RH (2001) Evaporites and siliciclastics of the
Permian Nippewalla Group, Kansas and Oklahoma: a case for
nonmarine deposition. Sedimentology 48:165–188
Benison KC, Bowen BB, Oboh-Ikuenobe FE, Jagniecki EA, LaClair
DA, Story SL, Mormile MR, Hong BY (2007) Sedimentol-
ogy of acid saline lakes in southern Western Australia: newly
described processes and products of an extreme environment. J
Sediment Res 77:366–388
Benison KC, Knapp JP, Dannenhoffer JM (2011) The Pennsylvanian
Pewamo Formation and associated Haybridge strata: toward the
resolution of the Jurassic Ionia red bed problem in the Michigan
Basin, USA. J Sediment Res 81:459–478
Benison KC, Zambito JJ, Soreghan GS, Soreghan MJ, Foster TM,
Kane MJ (2013) Permian red beds and evaporites of the Amoco
Rebecca K. Bounds core, Greeley County, Kansas: Implications
for science and industry. In: Dubois MK, Watney WL, Tollefson
(eds) Mid-continent core workshop: from source to reservoir
to seal. Mid-Continent Section American Association of Petro-
leum Geologists, Kansas Geological Survey and Kansas Geo-
logical Society, Wichita, Kansas, pp 9–14
Bloemendal J, Liu XM, Rolph TC (1995) Correlation of the magnetic
susceptibility stratigraphy of Chinese loess and the marine oxy-
gen isotope record: chronological and palaeoclimatic implica-
tions. Earth Planet Sci Lett 131:371–380
Brigham-Grette J, Haug GH, The Climate Working Group (2007) Cli-
mate dynamics and global environments: a community vision
for the next decade in ICDP. In: Harms UC, Koeberl C, Zoback
MD (eds) Continental scientific drilling: a decade of progress
and challenges for the future. Springer, Berlin, pp 53–94
Buijs GJA, Goldstein RH (2012) Sequence architecture and palaeocli-
mate controls on diagenesis related to subaerial exposure of ice-
house cyclic Pennsylvanian and Permian carbonates. Int Assoc
Sedimentol Spec Publ 45:55–80
Carter LS, Kelley SA, Blackwell DD, Naeser ND (1998) Heat flow
and thermal history of the Anadarko Basin, Oklahoma. Am
Assoc Pet Geol Bull 82:291–316
Chen J, Chen Y, Liu LW, Ji JF, Balsam W, Sun YB, Lu HY (2006)
Zr/Rb ratio in the Chinese loess sequences and its implication
for changes in the East Asian winter monsoon strength. Geo-
chim Cosmochim Acta 70:1471–1482
Clyde WC, Wing SL, Gingerich PD (2013) Bighorn Basin Coring
Project (BBCP): a continental perspective on early Paleogene
hyperthermals. Sci Drill 16:21–31
Cohen AS (2011) Scientific drilling and biological evolution in
ancient lakes: lessons learned and recommendations for the
future. Hydrobiol. doi:10.1007/s10750-010-0546-7
Cohen AS, Arrowsmith JR, Behrensmeyer AK, Campisano CJ, Fei-
bel CS, Fisseha S, Bedaso Z, Lockwood CA, Mbua E, Olago D,
Potts R, Reed K, Renaut R, Tiercelin JJ, Umer M (2009) Under-
standing paleoclimate and human evolution through the Homi-
nin Sites and Paleolakes Drilling Project. Sci Drill 8:60–65
Dean WE, Arthur MA (1998) Cretaceous Western Interior Seaway
drilling project: an overview. In: Dean WE, Arthur MA (eds)
Stratigraphy and paleoenvironments of the cretaceous western
interior seaway. USA SEPM Society for Sedimentary Geology
Concepts in Sedimentology and Paleontology 6:1–10
Dean WE, Arthur MA, Sageman BB, Lewan MD (1995) Core
descriptions and preliminary geochemical data for the Amoco
Production Company Rebecca K. Bounds #1 well, Greeley
County, Kansas. US Geol Surv. Open-File Rep. 95-209, 243 pp
Denison RE, Kirkland DW, Evans R (1998) Using strontium isotopes
to determine the age and origin of gypsum and anhydrite beds.
Geology 106:1–17
Dickinson WR, Gehrels GE (2009) Use of U–Pb ages of detrital zir-
cons to infer maximum depositional ages of strata: a test against
a Colorado Plateau database. Earth Planet Sci Lett 288:115–
125. doi:10.1016/j.epsl.2009.09.013
Ding ZL, Derbyshire E, Yang SL, Yu ZW, Xiong SF, Liu TS (2002)
Stacked 2.6-Ma grain size record from the Chinese loess based
on five sections and correlation with the deep-sea δ18O record.
Paleoceanography 17:1033
Dubiel RF, Smoot JP (1995) Criteria for interpreting paleoclimate
from red beds—a tool for Pangean reconstructions. In: Embry
AF, Beauchamp B, Glass DJ (eds) Pangea: global environments
and resources. Can Soc Pet Geol Mem 17:295–310
Dubois MK, Goldstein RH, Hasiotis ST (2012) Climate-controlled
aggradation and cyclicity of continental loessic siliciclas-
tic sediments in Asselian-Sakmarian cyclothems, Permian,
Hugoton embayment, USA. Sedimentology 59:1782–1816.
doi:10.1111/j.1365-3091.2012.01326.x
Evans JE, Reed JM (2007) Integrated loessite-paleokarst depositional
system, early Pennsylvanian Molas Formation, Paradox basin,
southwest Colorado, USA. Sediment Geol 195:161–181
Foster TM (2013) Environments and provenance of redbeds, of the
dog creek shale (midcontinent): implications for Middle Per-
mian Paleoclimate in western Pangaea. M.S. Thesis, University
of Oklahoma, Norman, Oklahoma, USA, 103 pp
Foster TM, Soreghan GS, Soreghan MJ, Benison KC, Elmore RD
(2014) Climatic and palaeogeographic significance of aeolian
sediment in the Middle Permian Dog Creek Shale (Midconti-
nent US). Palaeogeogr Palaeoclimatol Palaeoecology 402:12–
29. doi:10.1016/j.palaeo.2014.02.031
Frakes LA (1979) Climates through geologic time. Elsevier, Amster-
dam, 304 pp
Giles JM, Soreghan MJ, Benison KC, Soreghan GS, Hasiotis ST
(2013) Lakes, loess, and paleosols in the Permian Wellington
Formation of Oklahoma, USA: implications for paleoclimate
and paleogeography of the Midcontinent. J Sediment Res
83:825–846. doi:10.2110/jsr.2013.59
Golonka J, Ford DW (2000) Pangean (Late Carboniferous-Middle
Jurassic) paleoen-vironment and lithofacies. Palaeogeogr Pal-
aeoclimatol Palaeoecology 161:1–34
Golonka J, Ross MI, Scotese CR (1994) Phanerozoic paleogeographic
and paleoclimatic modeling maps. In: Embry AF, Beauchamp
B, Glass DJ (eds) Pangea: global environments and resources.
Can Soc Pet Geol Mem 17:47
Gradstein FM, Ogg JG, Schmitz M, Ogg G (2012) The
geologic time scale 2012. Elsevier. doi:10.1016/
B978-0-444-59425-9.00024-X
Ham WE (1960) Middle Permian evaporites in southwestern Okla-
homa. 21st Int Geol Congr 12:138–151
Hao QZ, Guo ZT (2005) Spatial variations of magnetic susceptibility
of Chinese loess for the last 600 kyr: implications for monsoon
evolution. J Geophys Res 110:B12101
Hatte C, Pessenda LC, Lang A, Paterne M (2001) Development of an
accurate and reliable 14C-chronology for loess sequences: appli-
cation to the loess sequence of Nubloch (Rhine Valley, Ger-
many). Radiocarbon 43:611–618
Heckel P (2008) Pennsylvanian cyclothems in Midcontinent
North America as far-field effects of waxing and waning of
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
Gondwana ice sheets. Geol Soc Am Spec Publ 441:275–289.
doi:10.1130/2008.2441(19
Hemmerich MJ, Kelley SA (2000) Patterns of Cenozoic denudation
on the Southern High Plains. Am Assoc Pet Geol Bull 84:1239
Hu X-F, Du Y, Guan C-L, Xue Y, Zhange G-L (2014) Color variations
of the Quaternary Red Clay in southern China and its paleocli-
matic implications. Sediment Geol 303:15–25
IODP (2008) Understanding climatic variability by scientific ocean
drilling. IODP thematic report series 1, 31 pp
Johnson KS (1978) Stratigraphy and mineral resources of Guada-
lupian and Ochoan rocks in the Texas panhandle and western
Oklahoma. N M Bur Mines Min Res Circ 159:57–62
Johnson KS (1989a) Geologic evolution of the Anadarko Basin. Okla
Geol Surv Circ 90:3–12
Johnson SY (1989b) Significance of loessite in the Maroon Formation
(Middle Pennsylvanian to Lower Permian), Eagle Basin, north-
western Colorado. J Sediment Pet 59:782–791
Kane MM (2013) Detrital zircon geochronology of a core in western
Kansas: implications for Permian paleogeography and paleocli-
mate. M.S. Thesis, University of Oklahoma, Norman, Okla-
homa, USA 123 pp
Kent DV, Muttoni G (2003) Mobility of Pangea: Implications for late
Paleozoic and early Mesozoic paleoclimate. In: LeTourneau
PM, Olsen PE (eds) The great rift valleys of Pangea in east-
ern North America, vol 1, tectonics, structure, and volcanism.
Columbia University Press, New York, pp 11–20
Kluth CF, Coney PJ (1981) Plate tectonics of the ancestral Rocky
Mountains. Geology 9:10–15
Kukla GJ, An ZS (1989) Loess stratigraphy in central China. Palaeo-
geogr Palaeoclimatol Palaeoecology 72:203–225
Kutzbach JE, Gallimore RG (1989) Pangean climates: megamonsoons
of the megacontinent. J Geophys Res 94:3341–3357
Lang A, Hatte C, Rousseau D-D, Antoine P, Fontugne MR, Zoller
L, Hambach U (2003) High-resolution chronologies for loess:
comparing AMS 14C and optical dating results. Quat Sci Rev
22:953–959
Liang L, Sun Y, Yao Z, Liu Y, Wu F (2012) Evaluation of high-resolu-
tion elemental analyses of Chinese loess deposits measured by
X-ray fluorescence core scanner. Catena 92:75–82
Liu TS (1985) Loess and environment. China Ocean Press, Beijing,
pp 1–106
Liu XM, Rolph T, Bloemendal J, Shaw J, Liu TS (1995) Quantita-
tive estimates of palaeoenvironment of palaeoprecipitation at
Xifeng, in the Loess Plateau of China. Palaeogeogr Palaeocli-
matol Palaeoecol 113:243–248
Liu TS, Ding ZL, Rutter N (1999) Comparison of Milankovitch peri-
ods between continental loess and deep sea records over the last
2.5 Ma. Quat Sci Rev 18:1205–1212
Loope DB, Steiner MB, Rowe CM, Lancaster N (2003) Tropical
westerlies over Pangaean sand seas. Sedimentology 51:315–322
Lowenstein TK, Hardie LA (1985) Criteria for the recognition of salt-
pan evaporites. Sedimentology 32:627–644
Lowenstein TK, Li J, Brown C, Roberts SM, Ku T-L, Luo S, Yang W
(1999) A 200 k.y. paleoclimate record from Death Valley salt
core. Geology 27:3–6
McKee ED, Oriel SS (1967) Paleotectonic investigations of the Per-
mian Systems in the United States. US Geol Surv Prof Publ
515, 268 pp
Melles M, Brigham-Grette J, Minyuk PS, Nowaczyk NR, Wennrich
V, DeConto RM, Anderson PM, Andreev AA, Coletti A, Cook
TL, Haltia-Hovi E, Kukkonen M, Lozhkin AV, Rosen P et al
(2012) 2.8 Million years of Arctic climate change from Lake
El’gygytgyn, NE Russia. Science 337:315–320
Myrow PM (1990) A new graph for understanding colors of mud
rocks and shales. J Geol Educ 8:16–20
National Research Council (NRC) (2011) Understanding Earth’s deep
past: lessons for our climate future. National Academies Press,
National Research Council, 212 pp
Norton GH (1939) Permian red beds of Kansas. Am Assoc Pet Geol
Bull 23:1751–1803
Oboh-Ikuenobe FE, Sanchez Botero CA (2013) Preservation of
organic-walled microfossils in red sediments: recovery of useful
palynomorph taxa in lacustrine sediments in Western Australia.
Geol Soc Am Ann Meet Abst. Denver, USA, pp 343–317
Olsen PE, Kent DV, Cornet B, Witte WK, Schlische RW (1996) High-
resolution stratigraphy of the Newark rift basin (early Meso-
zoic, eastern North America). Geol Soc Am Bull 108:40–77
Olsen PE, Kent DV, Geissman JW, Bachmann G, Blakey RC, Gehrels
GE, Irmis RB, Kuerschner W, Molina-Garza R, Mundil R, Sha
JG (2010) The Colorado Plateau Coring Project (CPCP): 100
million years of earth system history. Earth Sci Front 17:55–63
Parrish JT (1993) Climate of the supercontinent Pangea. J Geol
101:215–233
Parrish JT, Peterson F (1988) Wind directions predicted from global
circulation models and wind directions determined from eolian
sandstones of the western United States—a comparison. Sedi-
ment Geol 56:261–282
Roberts SM, Spencer RJ (1995) Paleotemperatures preserved in fluid
inclusions in halite. Geochim Cosmochim Acta 59:3929–3942.
doi:10.1016/0016-7037(95)00253-V
Robinson PL (1973) Paleoclimatology and continental drift. In: Tar-
ling DH, Runcorn SK (eds) Implications of continental drift to
the earth sciences I. Academic Press, London, pp 449–476
Ross CA, Ross JRP (1988) Late Paleozoic transgressive—regres-
sive deposition. In: Wilgus CK, Hastings BS, Kendall CGSC,
Posamentier HW, Ross CA, Van Wagoner JC (eds) Sea-level
changes—an integrated approach. SEPM Spec Publ 42:227–247
Ross CA, Ross JRP (1994) Permian sequence stratigraphy and fossil
zonation. In: Embry AF, Beauchamp B, Glass DJ (eds) Pangea:
global environments and resources. Can Soc Pet Geol Mem
17:219–231
Ross CA, Ross JRP (1995) Permian sequence stratigraphy. In: Scholle
PA, Peryt TM, Ulmer-Scholle DS (eds) Paleogeography, pale-
oclimate, stratigraphy, the Permian of Northern Pangea, vol 1.
Springer, Berlin, pp 98–123
Sánchez Botero CA, Oboh-Ikuenobe FE, Macphail MK (2013)
First fossil pollen record of the Northern Hemisphere species
Aglaoreidia cyclops Erdtman, 1960 in Australia. Alcheringa
37:1–5
Sawin RS, Franseen EK, West RR, Ludvigson GA, Watney WL
(2008) Clarification and changes in Permian stratigraphic
nomenclature in Kansas. Kans Geol Surv Bull 254:1–4
Scholz C, Johnson TC, Cohen AS, King JW, Peck JA, Overpeck JT,
Talbot MR et al (2007) East African mega-droughts between
135 and 75 thousand years ago and bearings on early-modern
human origins. Proc Natl Acad Sci USA 42:16416–16421. doi:
10.1073/pnas.0703874104
Scotese CR (1999) PALEOMAP animations, paleogeography, PALE-
OMAP project: Arlington. University of Texas at Arlington,
Department of Geology, Arlington
Sheldon ND (2005) Do red beds indicate paleoclimatic conditions?
A Permian case study. Palaeogeogr Palaeoclimatol Palaeoecol
228:305–319
Shen S-Z, Schneider JW, Angiolini L, Henderson CM (2013) The
international Permian timescale: March 2013 update. In: Lucas
SG et al (eds) The Carboniferous-Permian transition. New Mex-
ico Mus Nat Hist Sci Bull 60:411–416
Slingerland R, Furlong KP (1989) Geodynamic and geomorphic evo-
lution of the Permo-Triassic Appalachian mountains. Geomor-
phology 2:23–37
Author's personal copy
Int J Earth Sci (Geol Rundsch)
1 3
Soreghan G (1992) Preservation and paleoclimatic significance of
eolian dust in the Ancestral Rocky Mountains province. Geol-
ogy 20:1111–1114
Soreghan GS, Cohen AS (2013) Scientific drilling and the evolution of
the earth system: climate, biota, biogeochemistry and extreme
systems. Sci Drill 16:63–72. doi:10.5194/sd-16-63-2013
Soreghan MJ, Francus P (2004) Processing backscattered electron
digital images of thin section. In: Francus P (ed) Image analysis,
sediments and paleoenvironments. Developments in paleoenvi-
ronmental research, vol 7. Kluwer, Netherlands, pp 203–225
Soreghan G, Elmore R, Katz B, Cogoini M, Banerjee S (1997) Pedo-
genically enhanced magnetic susceptibility variations preserved
in Paleozoic loessite. Geology 25:1003–1006
Soreghan M, Soreghan G, Hamilton M (2002) Paleowinds inferred
from detrital-zircon geochronology of upper Paleozoic loessite,
western equatorial Pangea. Geology 30:695–698
Soreghan G, Soreghan M, Hamilton M (2008) Origin and significance
of loess in late Paleozoic western Pangaea: a record of tropi-
cal cold? Palaeogeogr Palaeoclimatol Palaeoecol 268:234–259.
doi:10.1016/j.palaeo.2008.03.050
Soreghan GS, Keller GR, Gilbert MC, Chase CG, Sweet D (2012)
Load-induced subsidence of the Ancestral Rocky Mountains
recorded by preservation of Permian landscapes. Geosphere
8:654–668. doi:10.1130/GES00681.S1
Soreghan MJ, Heavens N, Soreghan GS, Link PK, Hamilton MA
(2014) Abrupt and high-magnitude changes in atmospheric cir-
culation recorded in the Permian Maroon Formation, tropical
Pangaea. Geol Soc Am Bull. doi:10.1130/B30840.1
Steiner M (2006) The magnetic polarity time scale across the Per-
mian-Triassic boundary. Non-marine Permian biostratigraphy
and biochronology. Geol Soc Lond Spec Publ 265:15–38
Sun JM (2002) Provenance of loess material and formation of loess
deposits on the Chinese Loess Plateau. Earth Planet Sci Lett
203:845–859
Sur S, Soreghan G, Soreghan M, Yang W, Saller A (2010) A record
of glacial aridity and Milankovitch-scale fluctuations in atmos-
pheric dust from the Pennsylvanian tropics. J Sediment Res
80:1046–1067
Sweet AC, Soreghan GS, Sweet DE, Soreghan MJ, Madden
AS (2013) Permian dust in Oklahoma: source and origin
for Middle Permian (Flowerpot-Blaine) redbeds in West-
ern Tropical Pangaea. Sediment Geol 284–285:181–196.
doi:10.1016/j.sedgeo.2012.12.006
Swineford A (1955) Petrography of Upper Permian rocks in south-
central Kansas. Kans Geol Surv Bull 111:179
Tabor NJ, Montanez IP (2002) Shifts in late Palaeozoic atmospheric
circulation over western equatorial Pangaea: insights from
pedogenic mineral d18O compositions. Geology 30:1127–1130
Tabor NJ, Poulsen CJ (2008) Palaeoclimate across the Late Pennsyl-
vanian-Early Permian tropical palaeolatitudes: a review of cli-
mate indicators, their distribution, and relation to palaeophysio-
graphic climate factors. Palaeogeogr Palaeoclimatol Palaeoecol
268:293–310
Tierney JE (2010) An organic geochemical perspective on tropical
East African paleoclimate. PhD dissertation, Brown University,
Providence, Rhode Island, 237 pp
Tramp KL, Soreghan GS, Elmore RD (2004) Paleoclimatic inferences
from paleopedology and magnetism of the Permian Maroon
Formation loessite (Colorado, USA). Geol Soc Am Bull
116:671–686
Turner P (1980) Continental red beds. Concepts in sedimentology, vol
29. Elsevier, Amsterdam, 562 pp
Vandenberghe J, Lu HY, Sun DH, van Huissteden J, Konert M (2004)
The late Miocene and Pliocene climate in East Asia as recorded
by grain size and magnetic susceptibility of the Red Clay
deposits (Chinese Loess Plateau). Palaeogeogr Palaeoclimatol
Palaeoecol 204:239–255
Van der Voo B (1993) Paleomagnetism of the Atlantic, Tethys and
Iapetus Oceans. Cambridge University Press, Cambridge, p 411
Walker TR (1967) Formation of red beds in modern and ancient
deserts. Geol Soc Am Bull 78:353–368
Walker TR (1976) Diagenetic origin of continental red beds. In:
Faulke H (ed) The continental Permian in central, west, and
south Europe. D. Reidel Publishing, Netherlands, pp 240–282
Walker SH, Millhein KK (1989) An innovative approach to explora-
tion and exploitation drilling: the slim-hole high-speed drilling
system. In: Proceedings of the 64th annual technical conference
exhibition. SPE paper 19525, pp 73–89
West R, Miller K, Watney W (2010) The Permian System in Kansas.
Kans Geol Surv Bull 257:1–88
Zambito JJ, Benison KC (2013) Extremely high temperatures and
paleoclimate trends recorded in Permian ephemeral lake halite.
Geology 41:587–590
Benison KC, Zambito JJ, IV (2013) Correlation of red beds and evap-
orite units between surface and subsurface: addressing chal-
lenges for petroleum geology. Abst Prog Am Assoc Pet Geol
Ann Meet. AAPG Search and Discovery Article #90163
Zambito JJ, Benison KC, Foster T, Soreghan GS, Kane M, Soreghan
MJ (2012) Lithostratigraphy of the Permian red beds and evap-
orites in the Rebecca K. Bounds, Core, Greeley County, Kan-
sas. Kansas Geol Surv Open-File Rep 2012-15, 45 pp
Zhou LP, Oldfield F, Wintle AG, Robinson SG, Wang JT (1990)
Partly pedogenic origin of magnetic variations in Chinese loess.
Nature 346:737–739
Author's personal copy
... These altered outcrops do not provide accurate lithologies or stratigraphic thicknesses or relationships. The most complete data about ancient siliciclastic– evaporite rocks can be obtained through the study of cores from the subsurface (e.g., Soreghan et al. 2014). A prerequisite for these cores is that they be drilled with care to preserve halite. ...
... Coring the subsurface counterparts from the preservation depths window (, 1000–3000 feet; , 305–915 meters) with use of saltsaturated drilling fluids would allow well-preserved strata with excellent recovery, such as that of the Bounds core, to be obtained and used for a subsurface reference-section counterpart to the surface type section. A future continental drilling program that combines high-quality coring with nuclear logging would substantially benefit this effort to better understand mixed siliciclastic–evaporite units in the subsurface (Soreghan et al. 2014). ...
Article
Full-text available
Geologists tend to expect that rocks as they appear in outcrop are a very good approximation of the same strata in the subsurface. But this is not the case in mixed siliciclastic-evaporite units, strata that contain both a siliciclastic and an evaporite mineral component. Distinct differences in lithologic and stratigraphic descriptions of ancient silicicastic-evaporite strata exist between outcrop and core. These differences present challenges in stratigraphic nomenclature, lithologic correlation, age determinations, and interpretations about past depositional environments, diagenetic history, and paleoclimate. An example of siliciclastic-evaporite strata is the middle Permian to early Triassic gypsum- and/or anhydrite- (hereby referred to as gypsum/anhydrite) and halite-bearing red beds throughout the midcontinent of North America and equivalent strata from other Pangean deposits. Here, we use core and field observations of the middle Permian Nippewalla Group in Kansas to show the range of similarities and differences in lithologies and thicknesses of specific lithofacies in outcrop and at various depths. For example, the displacive halite lithology, consisting of red mudstones dominated by randomly oriented halite crystals, is abundant in cores and interpreted as saline mudflat deposits. The outcrop counterpart is less common, friable, massive red mudstones with some gypsum/anhydrite pseudomorphs after displacive halite crystals. We interpret these lithologic differences of this saline mudflat lithofacies as the result of excellent preservation of buried displacive halite and late-stage dissolution of near-surface halite. Additionally, bedded halite and intergranular halite cements in siltstones and sandstones are also vulnerable to late-stage dissolution near the surface, further modifying lithologies and thicknesses. Thickness of the Nippewalla Group in the Amoco Rebecca K. Bounds core of western Kansas is 931 feet (~ 284 meters) thick, but it is estimated to be only half as thick in the outcrop belt of south-central Kansas. To estimate amount of rock lost to late-stage dissolution, we conducted dissolution experiments on displacive halite units from cores, which resulted in 60 - 95% loss of rock thickness and mass due to dissolution of halite. The unconsolidated red sediment resulting from these dissolution experiments compares favorably with much of the fine massive red siliciclastics that make up the greater part of the Nippewalla Group outcrops. We propose that major lithologic differences resulting from near-surface diagenetic processes must be expected in any evaporite-bearing siliciclastic unit. Besides guiding correlation attempts between surface and subsurface, this knowledge is vital for making accurate interpretations of depositional and diagenetic history of siliciclastic-evaporite units.
... Magnetostratigraphic analysis of a 1,600-m continuous core drilled in 1988 by Amoco Production Company in western Kansas (Fig. 1) indicated that the end of the Kiaman Superchron, a long period of reversed polarity that extended from the late Carboniferous to part of the Guadalupian, was nearly coincident with the Dog Creek Shale-Whitehorse formational contact (Foster et al., 2014). Integration of these findings with stratigraphic and sedimentological studies of excellent Dog Creek, Blaine, and Flowerpot exposures in Blaine County, Oklahoma (Sweet et al., 2013;Foster et al., 2014;Soreghan et al., 2015) places the Dog Creek in the lower Wordian to upper Roadian, the Blaine in the upper half of the Roadian, and the Flowerpot almost entirely in the lower half of the Roadian, if we accept a mid-Wordian age for the Illawarra reversal. Although no paleomagnetic data constrain the lower part of this interval, one of the main Blaine County exposures, Cat Canyon ( Fig. 4, C), which was studied by Foster et al. (2014), is less than 5 km away from some of the Chickasha vertebrate sites, which are intercalated with the middle part of the Flowerpot Shale Formation. ...
Article
Full-text available
The youngest Paleozoic vertebrate-bearing continental deposits of North America are Middle Permian (Guadalupian) in age and occur in the Chickasha Formation (El Reno Group) of central Oklahoma and the lithostratigraphically lower San Angelo Formation (Pease River Group) of North-Central Texas. Although regarded originally as Guadalupian, these deposits have been assigned recently to the Early Permian on the basis of marine fossils and questionable lithostratigraphic extrapolations from marine to continental strata. A review of ammonoid genera recovered from the Blaine Formation, which overlies both the Chickasha and San Angelo in Oklahoma and Texas, shows that they range globally in age from the Early to Late Permian but most occur in the Guadalupian or Middle Permian. A modest but intensively studied paleobotanical record of compression fossils from the San Angelo, as well as palynomorphs in rocks associated with the Chickasha, presents an unquestionably Middle to Late Permian flora dominated by voltzian conifers. The Chickasha and San Angelo vertebrate assemblages are overwhelmingly dominated by large caseid synapsids and indicate a biostratigraphic signal of early Guadalupian. The occurrence of the tupilakosaurid temnospondyl Slaugenhopia, the parareptile Macroleter, and the eureptile Rothianiscus suggest a Roadian age (lowermost Guadalupian) given the global records of closely related forms. These plant and vertebrate assemblages contrast sharply with those of underlying Cisuralian rocks of the Hennessey Formation of Oklahoma and the Clear Fork Formation of Texas, both of which are much more fossiliferous than those of the Guadalupian in the region. A barren interval of up to 300 m in thickness separates these fossil-bearing intervals. This true void, first reported a half century ago by E.C. Olson, has not been recognized in recent biochronology studies. These findings, as well as those of other vertebrate paleontologists who have evaluated the San Angelo and Chickasha data by other means, strongly refute the notion of “Olson’s Gap” as currently entertained.
... Other cores in the region around Belfast have recorded no halite in the Belfast Group, but have described expulsion of brines upon drilling (McCann, 1990). This strongly suggests that the BHEF is not preserved in most drill cores due to dissolution of halite by dilute drilling fluids Soreghan et al., 2015). Thus, the Islandmagee-1 core, described herein, is exceptional in its excellent preservation of the BHEF. ...
Article
Full-text available
Ancient salt deposits preserve a record of highly specific environmental, climatic and biological conditions, including past surface water chemistry and water depths, local air temperatures, atmospheric composition, and halophilic microorganisms. This paper presents the first sedimentological study of the mid‐Late Permian Belfast Harbour Evaporite Formation of Northern Ireland. This formation is dominated by bedded halite, contains some siliciclastic mudstone and bedded anhydrite, and is cross‐cut by intrusive igneous rocks. The bedded halite lithology contains bottom‐growth chevron and cornet crystals, efflorescent crusts and dissolution pipes, which are evidence of a saline surface brine that underwent periods of evapoconcentration, desiccation and flooding. The mudstone lithology contains dewatering structures, intraclasts and mudcracks, which indicate flooding and desiccation in dry mudflats. Bedded anhydrite was deposited as beds of gypsum cumulate crystals in saline surface waters. The Belfast Harbour Evaporite Formation was formed by saline lakes with associated mudflats, based on sedimentary characteristics and supported by mineralogical, geochemical and stratigraphic context. Diagenetic features reflect dissolution and cementation at the surface and shallow subsurface in the depositional environment and limited late‐stage alterations. Syndepositional dissolution pipes in bedded halite were formed by flooding events. Early halite cement is present in dissolution pipes and mudstones. Gypsum/anhydrite repeatedly dehydrated and rehydrated to form an interlocking crystal mosaic. Later features include fluid migration and the intrusion of mafic rocks. The Belfast Harbour Evaporite Formation was deposited by an ephemeral saline lake and dry mudflat system in an arid climate. This study, when compared to age‐equivalent continental deposits elsewhere, suggests that arid settings with saline lakes existed across much of Pangea during the Permo‐Triassic. The Belfast Harbour Evaporite Formation stratigraphically underlies the extremely low pH saline lakes of the Mercia Mudstone Group, implying that it aids in understanding the of formation of Pangean acid brine lakes.
... A list of questions associated with the major and supplementary science topics appears in a data supplement table; overviews appear below. Coring is essential to (1) access buried upland records, (2) achieve continuous records in basinal regions, and (3) recover a pristine section suitable for application of various sedimentologic, geochemical, fluid-inclusion, rock-magnetic, magnetostratigraphic, paleontologic, and geochronologic approaches (i.e., Soreghan et al., 2015b;Benison et al., 2015). Additionally, during the course of both workshops, participants suggested considering expansion to a staged project that could incorporate coring in (1) multiple basins of western Europe and (2) in midpaleolatitude sites (e.g., China). ...
Article
Full-text available
Chamberlin and Salisbury's assessment of the Permian a century ago captured the essence of the period: it is an interval of extremes yet one sufficiently recent to have affected a biosphere with near-modern complexity. The events of the Permian – the orogenic episodes, massive biospheric turnovers, both icehouse and greenhouse antitheses, and Mars-analog lithofacies – boggle the imagination and present us with great opportunities to explore Earth system behavior. The ICDP-funded workshops dubbed “Deep Dust,” held in Oklahoma (USA) in March 2019 (67 participants from nine countries) and Paris (France) in January 2020 (33 participants from eight countries), focused on clarifying the scientific drivers and key sites for coring continuous sections of Permian continental (loess, lacustrine, and associated) strata that preserve high-resolution records. Combined, the two workshops hosted a total of 91 participants representing 14 countries, with broad expertise. Discussions at Deep Dust 1.0 (USA) focused on the primary research questions of paleoclimate, paleoenvironments, and paleoecology of icehouse collapse and the run-up to the Great Dying and both the modern and Permian deep microbial biosphere. Auxiliary science topics included tectonics, induced seismicity, geothermal energy, and planetary science. Deep Dust 1.0 also addressed site selection as well as scientific approaches, logistical challenges, and broader impacts and included a mid-workshop field trip to view the Permian of Oklahoma. Deep Dust 2.0 focused specifically on honing the European target. The Anadarko Basin (Oklahoma) and Paris Basin (France) represent the most promising initial targets to capture complete or near-complete stratigraphic coverage through continental successions that serve as reference points for western and eastern equatorial Pangaea.
... Although the term 'Danxia landform' is not as popular in other countries as in China, extensive international research has been conducted on continental red beds and the development of sandstone and conglomerate landforms [20][21][22][23][24][25]. Many spectacular red bed landscapes outside China have a similar appearance to Danxia landform, such as the 'rose-red' cliffs of Petra, Jordan [26]. ...
Article
Full-text available
As an erosional landform, the formation processes of Danxia landform are controlled by internal and external forces as well as lithologic properties. Using field data, we studied the role of lithologic properties on the formation of Danxia landform in Kongtongshan National Geopark, northwest China, through a series of experiments, including uniaxial compressive strength, identification analysis under polarizing microscope, X-ray diffraction analysis, inductively coupled plasma-mass spectrometry analysis, and scanning electron microscopy. The results show that the diagenesis degree, mineral composition, cement composition, degree of cementation, geochemical composition and element contents, and micro-structure influenced the structure and anti-weathering and anti-erosion abilities of the Danxia rock mass. Differential weathering of rock in different environments was an important force shaping the different types of Danxia landform. Weathering failure of the Danxia rock mass was the result of multiple combined factors; as well as lithology, other factors, such as those induced during tectonic uplift (i.e., faulting, jointing, and fracturing) and climate, cannot be neglected. Therefore, lithology played an important role in the structural development of Danxia landform, and different lithologies influenced its weathering rate and formation processes. Our findings can provide a reference for revealing the microscopic development of Danxia landform in arid and semi-arid areas. © 2019 Zhu et al. This is an open access article distributed under the terms of the Creative Commons Attribution License, which permits unrestricted use, distribution, and reproduction in any medium, provided the original author and source are credited.
... Radiometric dates from Bowring et al. (1998) and Nicklen 1865, related to the lithostratigraphy via sequence correlation of Rush & Kerans (2010). 1866Rebecca K Bounds core magnetostratigraphy from Soreghan et al. (2015), and additional details 1867 from Sawin et al. (2008). Buffels River composite (South Africa) from Tohver et al. (2015) Li & Wang (1989), Liu et 1881) in Yuan et al. (2014) & Meng et al. (2000. ...
Article
Full-text available
The reverse polarity Kiaman Superchron has strong evidence for at least three, or prob ably four, normal magnetochrons during the early Permian. Normal magnetochrons are during the early Asselian (base CI1r.1n at 297.94+0.33 Ma), late Artinskian (CI2n at 281.24+2.3 Ma), mid-Kungurian (CI3n at 275.86+2.0 Ma) and Roadian (CI3r.an at 269.54+1.6 Ma). The mixed-polarity Illawarra Superchron begins in the early Wordian at 266.66+0.76 Ma. The Wordian– Capitanian interval is biased to normal polarity, but the basal Wuchiapingian begins the beginning of a significant reverse polarity magnetochron LP0r, with an overlying mixed-polarity interval through the later Lopingian. No significant magnetostratigraphic data gaps exist in the Permian geomagnetic polarity record. The early Cisuralian magnetochrons are calibrated to a succession of fusulinid zones, the later Cisuralian and Guadalupian to a conodont and fusulinid biostratigraphy, and Lopingian magnetochrons to conodont zonations. Age calibration of the magnetochrons is obtained through a Bayesian approach using 35 radiometric dates, and 95% confidence intervals on the ages and chron durations are obtained. The dating control points are most numerous in the Gzhelian–Asselian, Wordian and Changhsingian intervals. This significant advance should provide a framework for better correlation and dating of the marine and non-marine Permian
Article
Full-text available
Siliciclastic strata of the Colorado Plateau attract attention for their striking red, green, bleached, and variegated colors that potentially record both early depositional and later diagenetic events. We investigated the proximal-most strata of the Paradox Basin, from their onlap contact with the Precambrian basement of the Uncompahgre Plateau to the younger Cutler strata exposed within 10 km of the Uncompahgre Plateau to attempt to understand the significance of the striking colors that occur here. These strata preserve a complex geology associated with buried paleorelief and sediment-related permeability variations at a major basin-uplift interface. Strata exposed within ∼1.5 km of the onlap contact exhibit a pervasive drab color in contrast to the generally red colors that predominate farther from this front. In-between, strata commonly host variegated red/green/bleached intercalations. Thin-section petrography, SEM, XRD, Raman spectroscopy, Mössbauer spectroscopy, and whole-rock geochemistry of samples representing different color variations from demonstrate that water–rock interactions charged the rocks with Fe(II) that persists primarily in the phyllosilicate fraction. Color variations reflect grain-size differences that allowed the reduction of fluids from regional fault and basement/fill contacts to permeate coarser-grained Cutler sediments. Hematite and chlorite occur in both red and green sediments but are absent in the bleached sediments. Pervasive hematite in both red and green layers suggests that sediments were hematite-rich before later alteration. Chlorite and smectite are elevated in green samples and inversely correlated with biotite content. Green coloration is generally associated with 1) coarser grain sizes, 2) spatial association with basement contacts, 3) elevated smectite and/or chlorite, 4) less total Fe but greater Fe(II)/Fe(III) primarily in the phyllosilicate fraction, and 5) uranium enrichment. The bleached coloration reflects the removal of pigmentary Fe(III) oxide, while the green coloration is due to the removal of pigmentary hematite and the abundance of Fe(II)-bearing phyllosilicates. Abundant mixed-layer and swelling clays such as smectite, illite/smectite, and chlorite/smectite (including tosudite) dominate the mineralogy of the clay fraction. These results are consistent with other studies demonstrating fault-associated fluid alteration in the Paradox Basin region. However, the pervasive greening was not observed in many of these studies and appears to reflect the unique aspects of the paleovalley system and the importance of biotite alteration to Fe(II)-bearing phyllosilicates.
Article
Full-text available
Cretaceous continental sediments in Sichuan Basin, China, have different colors, and the reasons for their formation are not determined. Based on mineralogical and geochemical characteristics, red beds and nonred beds in the Upper and Lower Cretaceous sedimentary strata in the western Sichuan Basin are described and tested in this study. The test and analysis of the mineral composition, element content, and iron speciation of mudstone samples with gray-green, gray, and red colors in Cangxi, Bailong, and Guankou formations found that the change in hematite content directly causes the color difference of samples. For red mudstone, the average chemical index of alteration, chemical index of weathering, weather eluviation index (Ba), Ca/(Mg*Al), and Al2O3/SiO2 index are 67.75, 79.94, 2.07, 0.26, and 0.26, respectively, indicating that chemical weathering is the most intense. The geochemical indexes corresponding to gray samples are 64.41, 74.91, 2.08, 0.19, and 0.24, respectively. Those corresponding to the gray-green samples are 62.30, 70.68, 2.17, 0.21, and 0.24, with the weakest chemical weathering. The ratio of Cu/Zn and the enrichment factor of V show that red and nonred bed samples are formed in weak oxidation and weak reduction environments, respectively. The red sample contains the highest content of hematite iron. The gray-green sample mainly represents paramagnetic ferrous in clay minerals. The geochemical contents of the gray sample’s three iron elements are slightly different, mainly trivalent iron. The change in iron speciation content in different color samples shows that the Fe element forming hematite in red bed samples may come from the weathering of source rock and clay minerals subjected to secondary weathering. At present, it is confirmed that different colors of samples are related to different weathering degrees of source rocks, which can be related to hot, dry/humid climates. It is necessary to distinguish the climate type in combination with other indicators.
Article
Full-text available
Given a phylogenetic tree that includes only extinct, or a mix of extinct and extant taxa, where at least some fossil data are available, we present a method to compute the distribution of the extinction time of a given set of taxa under the Fossilized-Birth-Death model. Our approach differs from the previous ones in that it takes into account (i) the possibility that the taxa or the clade considered may diversify before going extinct and (ii) the whole phylogenetic tree to estimate extinction times, whilst previous methods do not consider the diversification process and deal with each branch independently. Because of this, our method can estimate extinction times of lineages represented by a single fossil, provided that they belong to a clade that includes other fossil occurrences. We assess and compare our new approach with a standard previous one using simulated data. Results show that our method provides more accurate confidence intervals. This new approach is applied to the study of the extinction time of three Permo-Carboniferous synapsid taxa (Ophiacodontidae, Edaphosauridae, and Sphenacodontidae) that are thought to have disappeared toward the end of the Cisuralian (early Permian), or possibly shortly thereafter. The timing of extinctions of these three taxa and of their component lineages supports the idea that the biological crisis in the late Kungurian/early Roadian consisted of a progressive decline in biodiversity throughout the Kungurian.
Article
The Cenozoic Himalayan foreland basin sedimentary facies are widely studied by researchers as they preserve record of complex interplay between tectonics, sedimentation and climate. Palaeoclimate reconstructions from these sediments have been done using various proxy tools. Colour of sedimentary facies is considered as one of the diagnostic palaeoclimate/climate indicator and the red colour is commonly associated to be with hot, arid climate. The Himalayan foreland basin (HFB) sediments from the oldest to the youngest, comprise almost all red and grey shades and the bright red facies greatly fascinate Earth Scientists. These red to grey coloured Himalayan foreland basin sediments preserve a record of global to regional climatic cycles and tectonic pulses. Studies pertaining to red colour and climate, specifically its association with hot, arid climate, are overwhelming. In general, studies focusing on continental red beds with specific climate are contradictory and confusing till date. Paradoxically, we now know more about red colour and climate than about red color and tectonics. Therefore, representative Himalayan foreland basin red beds of different times in the NW Himalaya have been selected for the present study. The HFB red beds of Oligocene, middle Miocene and upper Pliocene age have been investigated to find out any specific relationship of tectonics and/or particular climate cycles with red colour of these sediments on the basis of compilation of morphological, micromorphological, sub-microscopic and geochemical studies. It has been found that the bright red and grey colour shades are reliably shown by finer facies, particularly palaeosols, and coarser facies of the HFB sediments respectively, whether formed in cool, humid conditions or warm, arid conditions. This suggests greater role of tectonics during foreland basin sedimentation for preservation of sediment colour. Also, the purpose of this study is to highlight the need for more such investigations at various temporal and spatial scales in foreland basin sediments so that the colour of sediment can be appropriately used in future investigations dealing with sediment colour as an indicator of specific climate.
Article
Full-text available
This paper outlines Permian nomenclature changes to Zeller (1968) that have been adopted by the Kansas Geological Survey. The Permian System/ Period, Cisuralian Series/Epoch, and Asselian Stage/Age are established at the base of the Bennett Shale Member of the Red Eagle Limestone. Series/epoch names Wolfcampian, Leonardian, and Guadalupian are retained and usage of Gearyan, Cimarronian, and Custerian is abandoned. The repositioned Carboniferous-Permian boundary divides the Council Grove Group into Carboniferous (Upper Pennsylvanian Series/Epoch; Virgilian Stage/Age) and Permian (Wolfcampian Series Epoch) segments.
Article
Full-text available
A US National Science Foundation-funded workshop occurred 17-19 May 2013 at the University of Oklahoma to stimulate research using continental scientific drilling to explore earth's sedimentary, paleobiological and biogeochemical record. Participants submitted 3-page "pre-proposals" to highlight projects that envisioned using drill-core studies to address scientific issues in paleobiology, paleoclimatology, stratigraphy and biogeochemistry, and to identify locations where key questions can best be addressed. The workshop was also intended to encourage US scientists to take advantage of the exceptional capacity of unweathered, continuous core records to answer important questions in the history of earth's sedimentary, biogeochemical and paleobiologic systems. Introductory talks on drilling and coring methods, plus best practices in core handling and curation, opened the workshop to enable all to understand the opportunities and challenges presented by scientific drilling. Participants worked in thematic breakout sessions to consider questions to be addressed using drill cores related to glacial-interglacial and icehouse-greenhouse transitions, records of evolutionary events and extinctions, records of major biogeochemical events in the oceans, reorganization of earth's atmosphere, Lagerstätte and exceptional fossil biota, records of vegetation-landscape change, and special sampling requirements, contamination, and coring tool concerns for paleobiology, geochemistry, geochronology, and stratigraphy-sedimentology studies. Closing discussions at the workshop focused on the role drilling can play in studying overarching science questions about the evolution of the earth system. The key theme, holding the most impact in terms of societal relevance, is understanding how climate transitions have driven biotic change, and the role of pristine, stratigraphically continuous cores in advancing our understanding of this linkage. Scientific drilling, and particularly drilling applied to continental targets, provides unique opportunities to obtain continuous and unaltered material for increasingly sophisticated analyses, tapping the entire geologic record (extending through the Archean), and probing the full dynamic range of climate change and its impact on biotic history.
Article
(A) Allegheny region, by Henry L. Berryhill, Jr (B) Gulf Coast region, by Eleanor J. Crosby (C) West Texas Permian basin region, by Steven S. Oriel, Donald A. Myers, and Eleanor J. Crosby (D) Northeastern New Mexico and Texas-Oklahoma Panhandles, by George H. Dixon (E) Oklahoma, by Marjorie E. MacLachlan (F) Central Midcontinent region, by Melville R. Mudge (G) Eastern Wyoming, eastern Montana, and the Dakotas, by Edwin K. Maughan (H) Middle Rocky Mountains and northeastern Great Basin, by Richard P. Sheldon, Earl R. Cressman, Thomas M. Cheney, and Vincent E. McKelvey (I) Western Colorado, southern Utah, and northwestern New Mexico, by Walter E. Hallgarth (J) Arizona and western New Mexico, by Edwin D. McKee (K) West Coast region, by Keith B. Ketner References cited (284pp) This professional paper is a supplementary volume to "Paleotectonic Maps of the Permian System" by McKee, Oriel, and others (1967), published by the U.S. Geological Survey as Miscellaneous Geologic Investigations Map I-450. The I-450 publication consists of 20 plates isopach and lithofacies maps, cross sections to accompany the maps, and interpretive and environmental maps and a summary of available geological information on each part of the Permian Period, an interpretation or reconstruction of Permian history, and brief discussions of environment, tectonics, and other significant features. The present volume explains and documents the maps and conclusions presented there. This study of the Permian System was made by 15 geologists, who were individually responsible for coverage of 18 regions. These authors are: Henry L. Berryhill, Jr. • Walter E. Hallgarth • Vincent E. McKelvey • Thomas M. Cheney • Keith B. Ketner • Melville R. Mudge • Earl R. Cressman • Marjorie E. MacLachlan • Donald A. Myers • Eleanor J. Crosby • Edwin K. Maughan • Steven S. Oriel • George H. Dixon • Edwin D. McKee • Richard P. Sheldon The Permian System of the 18 regions is described in 11 chapters. Each chapter presents an analysis of the basic data used, points out significant trends, and presents an interpretation, as well as alternative explanations where each occur, for the region concerned. The chapters and accompanying illustrations were coordinated and assembled by E. J. Crosby, E. D. McKee, W. W. Mallory, E. K. Maughan, and S. S. Oriel. Descriptive and documentary data are organized according to region, from east to west, and according to chronological sequence. Each chapter discusses, in order, rocks that underlie the Permian, the several intervals or divisions of the Permian (from oldest to youngest (table 1, in pocket)), and, finally, the rock units that directly overlie the Permian. Stratigraphic problems, the nature of contacts, trends in thickness and lithology, possible sources of sediment, environments of deposition, and paleotectonic implications of each interval are treated in that order.
Article
The book is a concise, integrated history of the climatic conditions of the Earth throughout all geolic time. Basin his interpretations on global reconstructions from rock magnetism and sea-floor spreading data, and using sedimentological, biological and chemical indicators of climate, the author documents and discusses the evolution of the Earth's climates on the world scale. New information on oceanic climates from DSDP is used in detailed reconstruction of the climatic changes of the Mesozoic and Cenozoic eras. An analysis of the thermal history of the Earth reveals a number of anomalies, the two most outstanding being the late Precambrian, with its widespread low-latitude glaciation, and the Mesozoic, when the ocean-atmosphere system retained an extraordinary amount of solar energy. -D.G.Tout
Chapter
The Cretaceous Western Interior Seaway Drilling Project was begun in 1991 under the auspices of the U.S. Continental Scientific Drilling Program. It was intended to be a multidisciplinary study of Cretaceous carbonate and siliciclastic rocks in cores from bore holes along a transect across the Cretaceous Western Interior Seaway. The study focuses on middle Cretaceous (Cenomanian to Campanian) strata that include, in ascending order, Graneros Shale, Greenhorn Formation, Carlile Shale, and Niobrara Formation. The transect includes cores from western Kansas, eastern Colorado, and eastern Utah. The rocks grade from pelagic carbonates containing organic-carbon-rich source rocks at the eastern end of the transect to nearshore coal-bearing units at the western end. These cores provide unweathered samples and the continuous depositional record required for geochemical, mineralogical, and biostratigraphic studies. The project combines biostratigraphic, paleoecological, geochemical, mineralogical, and high-resolution geophysical logging studies conducted by scientists from the U.S. Geological Survey, Amoco Production Company, and six universities.
Chapter
Approximately sixty transgressive-regressive depositional sequences are present in Carboniferous and Permian shallow-marine successions on the world’s stable cratonic shelves. These sequences were synchronous depositional events that resulted from eustatic sea-level changes. Based on currently available age correlations of rapidly evolved late Paleozoic tropical, subtropical, and temperate shelf faunas, the sequences on different cratonic shelves were time equivalent. These transgressive-regressive sequences averaged about 2 million years and ranged from 1.2 to 4.0 million years in duration. Local depositional conditions are important in controlling sedimentary patterns on different cratonic shelves. These conditions are affected by changes in sea level, strandline position, and drainage base level and are reflected in the sedimentary record. Because midsize sea-level fluctuations are usually widely identifiable in the stratigraphic record, they are useful aids in correlation. They are particularly helpful between regions that have contrasting depositional conditions, such as between a carbonate shelf starved of clastic sediments and a clastic-dominated shelf on which carbonates are rare. The appearance of new species and genera generally occurs above unconformities that signal new marine transgressive events and new depositional sequences. The durations of the hiatuses between these transgressive-regressive sequences are difficult to estimate. The hiatuses may represent cumulatively as much time, if not more, than the rock record. The numerous worldwide synchronous unconformities marking hiatuses of considerable duration within late Paleozoic shelf strata suggest that the fossil record may be very incomplete and preserves mostly biota that were extant during times of high sea level. Such an incomplete fossil record could easily be misinterpreted as a punctuated evolution having a highly irregular mutation rate.
Chapter
The Permian System contains a great number and diversity of depositional sequences (Fig. 1) which illustrate sedimentary responses to a series of sealevel fluctuations. These sea-level fluctuations had many different amplitudes and durations, and were accompanied by a wide spectrum of rates of deposition (Fig. 2). The mid-continent and southwestern North American stable cratonic successions serve as the basis for our Permian sea-level interpretations (Ross and Ross 1987a, b, 1988); however, equally useful sections appear to be present in China, particularly South China. In the southern hemisphere, Western Australia has marine and glacial-marine depositional sequences that may eventually help tie sea-level events in high and middle latitudes of Gondwana with those of low latitudes of cratonic North America and the Tethys.
Chapter
Convincing counterparts of classic ancient continental red beds, such as those of Permo-Triassic age in Europe and North Africa, and those of Pennsylvanian to Triassic age in western interior North America, are forming diagenetically today in deposits of Cenozoic age in the arid and semi-arid regions of southwestern United States and northwestern Mexico. Sequences of sediments that range from Holocene to mid-Tertiary age show critical early stages of reddening, and reveal that the pigment (hematite) originates from intrastratal alteration of iron-bearing minerals. Any type of iron-bearing mineral provides a potential source of iron for the pigment, but the most commonly occurring parent minerals are detrital ferromagnesian silicates such as olivine, augite, hornblende, biotite, etc., and detrital and authigenic iron-bearing clay minerals.