Rosenau Journal of Sedimentary Research 2013
Rosenau Journal of Sedimentary Research 2013
Rosenau Journal of Sedimentary Research 2013
net/publication/277376756
CITATIONS READS
50 621
4 authors, including:
Scott Elrick
University of Illinois, Urbana-Champaign
67 PUBLICATIONS 1,001 CITATIONS
SEE PROFILE
Some of the authors of this publication are also working on these related projects:
Sedimentology of the Fluvial Systems of the Clear Fork Formation in North-Central Texas: Implications for Early Permian Paleoclimate and Plant Fossil Taphonomy View
project
All content following this page was uploaded by Scott Elrick on 10 June 2015.
ABSTRACT: Distinct lateral and stratigraphic trends in paleosol morphology and clay mineralogy, in conjunction with the
stable-isotope composition of sub-millimeter-scale spherulitic siderite (sphaerosiderite) and flint-clay kaolinite in middle–upper
Pennsylvanian cyclic coal-bearing strata of the Illinois basin (IB) presented herein provide proxy records of middle–late
Pennsylvanian equatorial terrestrial environments. Collectively, these data provide a better understanding of the polygenetic
history of ancient soils preserved in cyclic strata and indicate that low-latitude Pennsylvanian IB soil profiles were influenced by
a combination of autogenic and allogenic controls.
Lateral variations in paleosol morphology from the interior of the basin to the northern margin indicate that soil development
was dominantly influenced by autogenic controls, such as differences in local paleohydrology (i.e., groundwater fluctuations)
and subsidence rates across the basin. Conversely, allogenic controls (i.e., glacioeustasy and climate) are interpreted to be
principally responsible for the preservation of features in gleyed Protosols, gleyed Vertisols, and gleyed calcic Vertisols that
occur in the interior of the IB. These paleosols are characterized by a polygenetic development history that includes (1) an
initial period of soil formation in well-drained environments with seasonal wetting and drying of the profile, followed by (2) a
subsequent period of waterlogging and development of reducing conditions.
Paleosols that lack evidence for extensive periods of waterlogging are restricted, with the exception of two examples, to the
northern margin of the basin, and they become abundant in the stratigraphic record near the middle–upper Pennsylvanian
boundary (uppermost Desmoinesian). The stratigraphic distribution of these paleosols, in conjunction with a decrease in the
relative abundance of kaolinite in the , 2 mm size fraction of IB paleosols near the Desmoinesian–Missourian (D-M)
boundary, is interpreted to reflect a shift in the overall low-latitude regional climate to drier conditions.
The stable-isotope compositions of sphaerosiderite and flint-clay kaolinite in middle–upper Pennsylvanian strata of the IB
provide an additional proxy record of past equatorial terrestrial environments and are used to constrain the magnitude of
paleoclimate change across the D-M (, Moscovian–Kasimovian) boundary. Sphaerosiderite d18OV-PDB and d13CV-PDB values
range from 23.6% to 20.5% and 212.8% to 23.2%, respectively. In particular, sphaerosiderite d18O values from the B
horizons of paleosols display an approximate 21.5% shift across the D-M boundary. This stratigraphically short isotopic shift
occurs across the D-M boundary, on the order of one cyclothem (, 400 kyr), and is followed by a subsequent shift on the order
of one cyclothem, back to more positive siderite d18O values.
The average d18OV-SMOW and dDV-SMOW values of kaolinite isolated from the , 0.2 mm size fraction of a latest
Desmoinesian flint-clay breccia are +21.0% and 246.5%. These values correspond to a latest Desmoinesian crystallization
temperature of 27 ± 2uC and a meteoric-water d18OV-SMOW value of 23.1%. Combination of sphaerosiderite d18O values and
co-occurring flint-clay kaolinite dD and d18O values suggest a possible warming of up to 6uC across the D-M boundary,
corresponding to an approximate temperature change of siderite crystallization from 27 to 33uC. This temperature increase, in
conjunction with distinct stratigraphic trends in the clay mineralogy and morphology of IB paleosols, is consistent with previous
paleoclimate interpretations, and it suggests that low-latitude Pennsylvanian paleoclimate became drier and warmer across the
D-M boundary.
Many studies consider glacioeustatic fluctuations associated with the environments, such as phyllosilicates (Savin and Epstein 1970; Lawrence
late Paleozoic Ice Age (LPIA) as the dominant mechanism responsible for and Taylor 1971, 1972; Bird and Chivas 1988, 1989; Lawrence and
the origin of Pennsylvanian cyclothems in Europe and North America Rashkes-Meaux 1993; Elliott et al. 1997; Stern et al. 1997; Vitali et al.
(Wanless and Shepard 1936; Heckel 1986, 1994; Langenheim and Nelson 2002; Tabor et al. 2002; Tabor and Montañez 2005), paleosol calcite
1992). The persistence of Pennsylvanian–Early Permian Euramerican (Cerling 1984; Quade et al. 1989; Cerling and Quade 1993; Tabor et al.
cyclothems has been used as evidence that cyclothems are primarily the 2004, 2006), and sphaerosiderite (Ludvigson et al. 1998; White et al. 2001;
result of repeated large-scale (. 100 m), high-frequency (104 to 105 kyr) Ufnar et al. 2001, 2002, 2004; Suarez et al. 2010) may provide important
eustatic sea-level fluctuations at low latitudes related to large-scale, short- information related to weathering and environmental conditions at the
term waxing and waning of Gondwanan ice sheets (Heckel 1977, 1986, time of formation in the early burial environment, such as the d18O values
1994, 2002; West et al. 1997; Soreghan 1994; Soreghan and Giles 1999; of local meteoric water and temperature of crystallization. In this regard,
Isbell et al. 2003; Haq and Schutter 2008). Yet, studies of low-latitude pedogenic minerals from paleosols, such as kaolinite and sphaerosiderite,
carbonate and clastic successions (Feldman et al. 2005; Bishop et al. 2009, have the ability to be potentially powerful proxies of terrestrial
2010) have documented much lower-magnitude (, 20 m) sea-level paleoenvironmental conditions, provided that the original isotopic
fluctuations associated with the LPIA (Heckel 2008; Rygel et al. 2008). compositions of these minerals have not been altered subsequent to
Other workers have suggested that the role of climate change is the most pedogenesis.
important control on the lithostratigraphic content of cyclothems. The purpose of this contribution is to (1) document the complex,
Specifically, glacial–interglacial-scale (104 to 105 kyr) periods of relative polygenetic history among 97 paleosol profiles preserved in Pennsylva-
wetness and dryness, in combination with variations in sediment influx, nian cyclothems, the detailed morphology and mineralogy of which were
are chiefly responsible for producing characteristic rock types in tropical presented in a companion paper (Part I; Rosenau et al. 2013), (2) provide
and subtropical basins in spite of global base-level changes associated a record of low-latitude climate change across sea-level lowstands, as
with glacioeustasy (Ziegler et al. 1987; Cecil 1990; Cecil et al. 1992, 1993; inferred from the stable-isotope composition of siderite cements (d18O,
West et al. 1997). These studies indicate that variations in ice volume d13C) and a coeval kaolinite from a flint-clay deposit (dD, d18O), and (3)
appear to be closely linked to changes in climate, and they demonstrate a discuss the spatial and temporal changes in paleohydrology as they relate
close correspondence between low-latitude tropical climate change and to interrelated regional (i.e., groundwater fluctuations) to global (i.e.,
high-latitude glacial–interglacial episodes in Gondwanaland (Montañez et climate change and glacioeustasy) controls. These observations are
al. 2007 and references therein; Poulsen et al. 2007; Peyser and Poulsen discussed in the context of proposed glacioeustatic and climate-driven
2008; Horton et al. 2012). Climatic changes in the paleotropics have also cyclothem models in an effort to fully address the development of
been attributed to other factors, such as increased monsoonal precipitation Pennsylvanian paleosols in the IB.
on western equatorial Pangaea (Parrish 1993; Soreghan et al. 2002a, 2002b;
Tabor et al. 2002; Tabor and Montañez 2002, 2004), continental drift across METHODS
climate belts (Gibbs et al. 2002; Tabor et al. 2008), reduced moisture sources
Siderite
on Pangaea (Tabor and Montañez 2002; Ziegler et al. 2002), or the
emplacement of orographic barriers due to the uplift of the central Pangean Forty-five siderite-bearing horizons from paleosols and associated
mountains (CPMs; Rowley et al. 1985; Soreghan 2007; Tabor and Poulsen sedimentary strata were sampled and sectioned to produce polished thin
2008 and references therein). These studies highlight the need for elucidating sections for petrographic and stable isotope analyses. Siderites from
the short-term and long-term controls responsible for the development of pedogenic units representing four distinctly different paleosol morphol-
paleoequatorial cylothems and associated paleoclimate changes during the ogies (Rosenau et al. 2013) were taken from five paleosol profiles in the
LPIA, as well as the importance of delineating the relations between MAC and ELY cores (Fig. 1). Following the classification scheme of
competing, and possibly interrelated, allocyclic and autocyclic processes Mack et al. (1993), these paleosols are gleyed Protosols, gleyed Vertisols,
that led to cyclothem development. In order to better understand the global gleyed vertic Calcisols, and vertic Calcisols (see fig. 4 in Rosenau et al.
effects of the LPIA, and its influence on Pennsylvanian paleoequatorial 2013). Samples from non-pedogenic units were taken from interlaminated
environments in the IB in particular, it is important to integrate climate siltstones and sandstones underlying the five paleosol profiles.
change across and within individual terrestrial successions in Pennsylvanian Screening techniques, as outlined below, were employed to confirm the
cyclothems, while considering also the competing, and possibly interrelated, presence and morphology of all mineral phases in the samples and for
effects of local, regional, and global processes. designation of samples suitable for stable carbon and oxygen isotope
Paleosols are widely recognized throughout the sedimentary record and analysis of siderite. At least two isotopic analyses of siderite from each
are commonly used to help reconstruct the geomorphology of ancient pedogenic or sedimentary horizon were obtained from each sample.
terrains and regional paleoenvironments and paleoclimates. This reflects Siderite-bearing horizons were examined and described from petrograph-
the observation that regional climate and hydrology are important ic thin sections under plane-polarized and cross-polarized light. Carbon-
controls on physical and chemical processes that affect soil morphological ate cement textures and fabrics were described in order to help distinguish
and mineralogical development (Birkeland 1984; Folkoff and Meente- between primary pedogenic and secondary (burial) carbonate mineral
meyer 1987; Yemane et al. 1996; Wilson 1999). As such, the stratigraphic phases. In particular, samples containing well-preserved sphaerosiderite,
and spatial distribution of soil morphology and clay mineralogy is as indicated by smooth margins and radial-concentric crystalline
commonly used in paleosol studies as an indicator of paleoclimate, microstructures (Ludvigson et al. 1998), are considered to be pedogenic.
paleoecology, paleohydrology, and ancient landscape evolution (Retal- Samples containing sphaerosiderite were processed further in order to
lack 1983; Joeckel 1991, 1995; Kraus and Bown 1993; Bestland et al. isolate them from other mineral phases.
1997; Stern et al. 1997; Tabor et al. 2002) and has provided a better After petrographic analysis, nine horizons were identified that
understanding of the allocyclic and autocyclic processes responsible for contained sphaerosiderite. Two techniques were used to further isolate
sediment accumulation and paleosol development in ancient sedimentary the sphaerosiderite from these samples for isotopic analysis: (1) drilling
systems (Kraus 1987; Wright and Robinson 1988; Smith 1990; Kraus and the carbonate sample from the corresponding slab of thin sections using a
Aslan 1993; McCarthy et al. 1999; McCarthy and Plint 2003). hand-held Dremel rotary tool with a diamond-carbide bit to collect
Additionally, many studies have demonstrated that the isotopic , 100 mg of sample powder for isotopic analysis or (2) isolating
composition of minerals that form in low-temperature weathering sphaerosiderite crystals from bulk matrix associated with the horizon
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 639
from which the thin section was made. The second technique is described of calcite, and rule out the presence of other carbonate species (e.g.,
in more detail in the following section. ankerite and dolomite) that may confound accurate stable carbon and
Isolation of Sphaerosiderite from Bulk Paleosol Matrix.—Due to oxygen isotopic analysis of siderite.
sampling constraints and the very fine size (# 1 mm) of the pedogenic
sphaerosiderites, a method was devised to separate the sphaerosiderites
Flint-Clay Kaolinite
from the bulk paleosol matrix by ultrasonic agitation in deionized water.
After disaggregation, the solution was passed through a stacked series of A kaolinite-dominated flint clay composed of breccia clasts in a
250 mm, 125 mm, 88 mm, and 44 mm sieves in order to isolate the different claystone breccia in the MAC core (Fig. 1), occurring at 79.8–82.1 m, was
size fractions. Heavy-liquid separation using lithium metatungstate analyzed in order to determine its stable oxygen and hydrogen isotope
solution (r 5 2.9 g/cc) was then employed to concentrate the relatively composition. Bulk matrix from the flint-clay clasts was disaggregated by
dense sphaerosiderite from less dense siliciclastics in the bulk matrix. The ultrasonic agitation in deionized water and the , 125 mm, , 2 mm, and
dense fraction from the 44–88 mm size fraction was then ground in a , 0.2 mm equivalent spherical diameter (e.s.d.) size fractions were
corundum mortar and pestle and , 200 mg of sample was treated with isolated from the matrix by centrifugation. The mineralogy of the
10% acetic acid solution for a minimum of 5 hours at room temperature , 125 mm size fraction was determined by X-ray diffraction (XRD) of
(, 23uC) in order to remove admixed calcite and ensure complete random powder mounts. Step-scan analyses were performed over a range
isolation of sphaerosiderite. In-house experiments on various synthetic of 2 to 70 u2h, a step size of 0.05 u2h, and 1 second count time per step.
mixtures of calcium carbonate (Carrara Marble) and siderite (Ward9s Each aliquot of the , 2 mm and , 0.2 mm size fractions was prepared on
Siderite) suggest that this concentration of acetic acid and reaction time is filter membranes and transferred to glass slides as oriented aggregates. A
sufficient to remove admixed calcium carbonate without affecting the set of chemical and heat treatments were performed on different aliquots
stability of the siderite (Fig. 2). Sphaerosiderite samples were then rinsed of the size fractions, including (1) K+ saturation at room temperature, (2)
five times with deionized water to remove excess acetic acid and raise the Mg2+ saturation at room temperature, (3) Mg2+ saturation and glycerol
pH to that of deionized water (, 5.5) and then dried overnight in an oven solvation at room temperature, and (4) K+ saturation and heating at
set at 50uC. Mineralogical identification of the 44–88 mm size fraction was 500uC for at least 2 hours. Step-scan analyses of each treatment were
determined from powder mounts using a Rigaku Ultima III X-ray performed over a range of 2 to 30 u2h, a step size of 0.04 u2h, and 1 second
diffractometer with Cu-Ka radiation over a range of 20 to 60u 2h with a count time per step. Mineralogical identification from XRD spectra
step size of 0.04 u2h and a 1 second count time to confirm the presence of follows the procedures outlined in Moore and Reynolds (1997). The
siderite, determine its proportion in the sample, ensure complete removal , 0.2 mm size fraction was used for isotopic analysis because it is
640 N.A. ROSENAU ET AL. JSR
RESULTS
r
FIG. 3.—Generalized stratigraphic column of the Pennsylvanian subsystem in
Illinois. Selected members are displayed for points of reference. The larger, detailed
stratigraphic column shows units that are discussed in this contribution (i.e., upper
Tradewater –lower Mattoon formations). The smaller, less-detailed stratigraphic
column shows all of the Pennsylvanian formations present in the Illinois basin.
Global stage-boundary date and North American series dates are based on North
American cyclothem calibration of Heckel (2008) in conjunction with radiometric
(U-Pb) dates of zircons from the Donets Basin of Eastern Europe (Davydov et al.
2010). Figure modified from Willman et al. (1975) and Nelson et al. (2011).
TABLE 1.— Summary of diagnostic features, stratigaphic and lateral distribution, and environmental interpretation of paleosol types in the Illinois basin.a 642
Dominant
Number Classification Diagnostic , 2 mm clay Dominant Environmental Stratigraphic
Pedotype of Occurrences (Mack et al. 1993) Features mineralogyb Pedogenic Processes Interpretation Interbasinal Variability Distribution
A 68 Histosol Abundant I, K, Ch Accumulation of Poorly-drained; humid, Coal thickness varies from , 1 cm to Tradewater Fm.,
degraded plant plant matter everwet 230 cm across the basin. Carbondale Fm.,
matter Shelburn Fm.,
Patoka Fm., Bond
Fm., Mattoon Fm.
B 31 Protosol Weakly developed I, K,Ch Intermittent, rapid Indeterminate Extent of redoximorphic features varies Tradewater Fm.,
horizonation, inputs of sediment; depending on paleolandscape position in Carbondale Fm.,
abundant fossil rate of sedimentation the basin. Paleosols from the northern Shelburn Fm.,
plant matter, greater than rate of margin display high-chroma (red) hues and Patoka Fm., Bond
mottling pedogenesis abundant fine to coarse yellow, red, and Fm., Mattoon Fm.
bluish-gray mottles, while those of the
interior of the basin are dominantly drab
(gley) color, with rare, fine to medium,
high-chroma mottles as well as abundant
sphaerosiderite and pyrite.
C 31 gleyed Vertisol Slickensides, I/S, K,I, Ch Shrink-swell (1) During active pedogenesis, Extent of redoximorphic features varies Tradewater Fm.,
wedge-shaped moderate to well drained depending on paleolandscape position in Carbondale Fm.,
aggregates, with seasonal variations in the basin. Paleosols from the northern Shelburn Fm.,
extensive soil moisture, followed by margin display high-chroma (red) hues and Patoka Fm., Bond
mottling (2) post-pedogenic abundant fine to coarse yellow, red, and Fm., Mattoon Fm.
alteration under water- bluish-gray mottles, while those of the
saturated, anoxic interior of the basin are dominantly drab
conditions (gley) color, with rare, fine to medium,
high-chroma mottles as well as abundant
sphaerosiderite and pyrite.
D 13 gleyed calcic Vertisol Slickensides, I/S, I, K Shrink-swell; (1) During active pedogensis, - Tradewater Fm.,
wedge-shaped pedogenic calcite well drained with seasonal Carbondale Fm.,
N.A. ROSENAU ET AL.
conditions
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 643
Stratigraphic
Distribution
and Roucoux 2010). With uncommon exceptions, coal layers overlie
paleosols (underclay) and are abruptly overlain by strata that were
deposited under marine or tidally influenced (tidal rhythmites, sensu
would have needed to prevail for several thousand years to form peat
thick enough to become commercial coal. If sea level rose too high, the
plants would drown and no peat would form, whereas if sea level were too
low, accumulated peat would be lost due to oxidation. The range of
optimal sea level could not fluctuate by more than a meter or two during
several thousand years. To maintain these optimal conditions across a
Well-drained; subhumid to
large area of the IB long enough to form even one extensive peat deposit
Environmental
Interpretation
we propose that most, if not all, of the peat deposits developed when the
paleo–water table was at a maximum during the lowstand systems tract
(LST; glacial intervals), in response to increased tropical rainfall brought
about by the development of high-pressure zones over Gondwanan ice
TABLE 1.— Continued.
precipitation
Shrink-swell;
I/S, K, I
and the Energy Shale above the Herrin Coal (Fig. 3, Fig. 7, MAC core at
, 102 m) are good examples. Both gray shales overlie thick, low-sulfur
Detailed descriptions of pedotypes can be found in Rosenau et al. (2013).
calcite, minor
aggregates,
Slickensides,
pedogenic
ended peat accumulation, and fine sediment rapidly buried the peat.
Superbly preserved plant fossils, including tree stumps in growth position,
attest to rapid burial (DiMichele and Nelson 1989; DiMichele et al. 1996;
DiMichele et al. 2007) and sparse invertebrate fossils signify fresh to
of Occurrences (Mack et al. 1993)
calcic Vertisol
Classification
(Fig. 11, Unit 5). In most cases black shale directly overlies coal or gray
b
a
LSC cores). A thin, basal limestone is locally present at the base of the Fossil community successions in Oklahoma support the claim that
black shale, commonly reduced to a layer of pyritized, fragmentary black shales were deposited in deep water, at or near highstand
brachiopod and bivalve shells. The contact with the underlying coal or, (Boardman et al. 1984). However, Bisnett and Heckel (1996) acknowl-
where present, the gray shale, is commonly sharp and erosional (Bauer edged that both near-shore and deep-water varieties of black shale exist,
and DeMaris 1982). Lacking fossils of benthic organisms, the black, and that in some cases (e.g., Anna Shale; Figs. 3, 6, , 70 m in the LSC
fissile shale is interpreted to have accumulated under anoxic conditions core) the unit progressed from shallow to deep water. We interpret the
(Heckel 1977). Zangerl and Richardson (1963) favored deposition of the black shale as being deposited during the transgressive systems tract
black shale in shallow water with a floating algal mat, whereas Heckel (TST; Fig. 11). Regional correlations show that black shales typically
(1977, 2002) envisioned black shale forming at water depths of . 100 m. pinch out near basin margins in the IB, and onto tectonic highs,
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 645
FIG. 5.—Measured paleosol-bearing sections of the upper part of the middle Pennsylvanian Tradewater Formation. Measurements depicted on core graphic logs are in
meters (bold text) and feet (italicized text) as measured from the top of the core. Black and white bar scale is in increments of 10 feet. Capital letters refer to paleosols of a
given pedotype. Encircled capital letters refer to pedotype discussed in Rosenau et al. (2013). Encircled numbers and acronyms above measured sections represent site
identification (Fig. 1). Positions of sphaerosiderite samples are denoted by red stars. Correlations of sections follow Willman et al. (1975) and Nelson et al. (2011).
Tentative correlations of stratigraphic units are denoted with a question mark (?).
whereas the overlying limestone persists farther paleo-landward (Nelson Limestone at , 346 m in the ELY core; Fig. 11, Unit 6). Most examples
et al. 2010). are fossiliferous lime mudstones, wackestones, and packstones containing
a diverse open-marine fauna including brachiopods, bryozoans, echino-
Marine Limestone (TST to HST).—Marine limestone overlies black derms, bivalves, gastropods, foraminifera, and phylloid algae. These filter
shale in complete, ‘‘idealized’’ IB cyclothems (Figs. 3, 6, St. David feeders favor water of close to normal salinity without a large amount of
646 N.A. ROSENAU ET AL. JSR
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 647
suspended sediment. Normal circulation of oxygenated water had to be (Heckel 1994; Feldman et al. 2005). This creates the dilemma previously
restored at a time when the depositional sites were beyond the reach of raised with respect to the timing of peat formation. Alternatively, we
significant terrigenous sedimentation. The greater geographic extent of interpret that soil formation and valley incisions occurred during the
marine limestones as compared to black phosphatic shales in the IB is regressive to early lowstand systems tracts (RST to LST; Fig. 11; Plint et
evidence that the limestone units represent late transgressive to highstand al. 2001).
deposits (Nelson et al. 2010) and therefore were deposited during the late
transgressive systems tract to highstand systems tract (TST to HST; Flint-Clay Breccia
Fig. 11).
Flint clay is a fine-grained, massive rock that breaks with conchoidal
fracture and contains an abundance of well-crystallized kaolinite (Keller
Deltaic Clastics (HST to RST).—In the IB, marine limestones generally
1968, 1975, 1981; Smith and O’Brien 1965). It has been described from
give way to an upward-coarsening succession of clastic rocks (Fig. 11,
lower through upper Pennsylvanian strata in the Illinois and Appalachian
Unit 7). This commonly begins with dark gray, siderite-bearing shale that
basins (Foose 1944; Waage 1950; Smith and O’Brien 1965; Williams 1960;
may contain a limited brackish-water to normal marine fauna. Dark gray
Williams et al. 1968) and is interpreted to form by chemical leaching and
shales grade upward to lighter gray silty shales, siltstones, and fine-
alteration of fine-grained sediments in low-lying acid swamps (Patterson
grained sandstones (Figs. 3, 5, , 338 m to , 345 m in the ELY core,
and Hosterman 1960; Keller 1968, 1975, 1981; Smith and O’Brien 1965).
Fig. 11). These rocks are commonly interlaminated and display a variety An uppermost Desmoinesian (upper Shelburn Formation) example was
of trace fossils along with tidal rhythmites. As reported by Devera (1989) encountered in the MAC core, where it occurs as clasts in a breccia
in slightly older rocks, diverse trace fossils that occur in these (Figs. 7, 12A, B). A detailed description of the deposit is presented here.
rhythmically laminated sediments display marine affinities. In most The flint clay occurs in a 230-cm-thick pebble–cobble, alternating mud-
cases, it appears the deltas filled nearly all of the available accommoda- and matrix-supported breccia, the top of which is located five meters
tion space. In a few cases, marine limestone occurs at, or near, the top of a above the Danville Coal (Figs. 7, FCB at , 79.8–82.1 m; Fig. 12A–C).
deltaic package. These ‘‘delta-top limestone’’ units are irregular in The deposit truncates an underlying Type C paleosol (gleyed Vertisol).
distribution, mostly limited to areas of a few square kilometers or tens of The lower 43 cm of the unit contains successive layers that include the
kilometers. The Carlinville (Patoka Formation) and Bunje (Bond following: (1) 3–4 cm angular clasts of massive, fine-grained flint clay, (2)
Formation) Limestone members are examples (Fig. 3). We interpret the 4–8 cm wavy laminated siltstone and sandstone clasts, (3) 4–9 cm-long
deltaic clastic deposits as products of rapidly prograding deltas that were massive intermixed sphaerosiderite (75–100 mm in diameter) and
deposited during the late highstand to early regressive systems tracts rhombohedral crystalline dolomite (, 225 mm long 3 , 90 mm wide;
(HST to RST; Fig. 11). Fig. 12D, E), (4) thin organic-rich shale clasts up to 3–4 cm long and 0.5–
1 cm wide, (5) rare clasts of coal # 1 cm in diameter, and (6) 7-cm-
Soil Formation and Valley Incision (RST to LST).—Because highstand long 3 4-cm-wide variegated, red and green clay clasts crosscut by
deltas filled nearly all the available space in the IB, the sea bottom became sphaerosiderite-cemented channels and fractures (Fig. 12A–C, F). The
exposed during the early stages of the regression that marked the advent middle 130 cm (43–173 cm) is a mud-supported breccia with cobble-sized
of the next glacial episode. This set in motion the processes of subaerial flint-clay clasts with fractures filled by sphaerosiderite (75–100 mm in
weathering (soil formation) upon interfluves during fluvial incision diameter; Fig. 12A–C, F), , 1-cm-long black shale clasts in a pale yellow
(Fig. 11, Unit 2). Deltaic distributary channels and incised-valley-fill (2.5Y 8/2) clay-rich matrix, common laminated greenish-gray (G1
sequences are a prominent part of the rock record in the IB (Figs. 3, 11). 5/10GY) siltstone clasts, and abundant sphaerosiderite (75–100 mm in
Some of the smaller channels in the deltaic packages undoubtedly diameter) evenly dispersed throughout.
represent deltaic distributary channels. These sequences (1) feature The upper 57 cm is dominantly a mixture of siltstone and mudstone
sandstone filling, (2) are of moderate width (tens to a few hundreds of with (1) abundant pebble-size flint-clay clasts that are fractured and filled
meters) and areal extent, (3) exhibit moderate depths of incision (5 to with sphaerosiderite (75–100 mm in diameter) and rhombohedral dolomite
15 meters, and never below the base of the deltaic package), and (4) lack (, 225 mm long 3 , 90 mm wide; Fig. 12D, E), (2) abundant siderite-
soil formation on the upper surface (Fig. 11). Channels that result from cemented vertical fractures up to 0.5 cm wide and 8 cm long (Fig. 12A–
valley incision as sea level dropped are generally much larger (kilometers C), (3) common medium to coarse micritic carbonate clasts that are
wide). They cut through underlying coal units and commonly cut through crosscut by sphaerosiderite (75–100 mm in diameter) and rhombohedral
two or more underlying cyclothems (e.g., Crown Mine Sandstone, Fig. 3). dolomite (, 225 mm long 3 , 90 mm wide; Fig. 12D, E), and (4)
Some channels have been mapped across the entire IB (e.g., Hopkins abundant disorthic siderite, crystalline dolomite, sparry calcite, and
1958; Andresen 1961; Potter 1962, 1963). As mapped, these channels vary sphaerosiderite clasts (Fig. 12F). The upper boundary of this unit is an
from nearly linear to strongly sinuous or meandering. Fill is largely erosional surface that is scoured by the overlying micaceous, greenish-
sandstone in the lower part, with planar cross bedding, signifying fluvial gray siderite-rich siltstone. The , 125 mm size fraction of the flint-clay
channel processes. The upper part of the valley-fill sequence may be finer clasts from the lower 43 cm of the breccia are composed of a mineral
grained and show tidal influence (Fig. 11). Some previous authors assemblage dominated by kaolinite and quartz (Fig. 13A). The , 2 mm
attributed the soil formation and valley cutting to lowstand (maximum fraction is a mixture of kaolinite and quartz, whereas the , 0.2 mm size
glaciation), with limestone and delta progradation during regression fraction is pure kaolinite (Fig. 13B, C).
r
FIG. 6.—Measured paleosol-bearing sections of the middle Pennsylvanian Carbondale Formation. Measurements depicted on core graphic logs are in meters (bold
text) and feet (italicized text) as measured from the top of the core. Black and white bar scale is in increments of 10 feet. Wavy lines to the left of measured sections define
tops of paleosols. Capital letters refer to paleosols of a given pedotype. Encircled capital letters refer to pedotype discussed in Rosenau et al. (2013). Encircled numbers
and acronyms above measured sections represent site identification (Fig. 1). Positions of micritic siderite samples are denoted by black stars. Correlations of sections
follow Willman et al. (1975) and Nelson et al. (2011).
648 N.A. ROSENAU ET AL. JSR
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 649
Siderite-bearing horizons display two different textures: (1) micritic Sphaerosiderite and Kaolinite Stable Isotopes as
nodules of various size and morphology and (2) sub-millimeter-scale Paleoenvironmental Proxies
spherules composed of radial-fibrous crystals (sphaerosiderite; Ludvigson As discussed, two distinct types of siderite are recognized in this study:
et al. 1998). A complete description of the occurrence and micromor- (1) micritic cements associated with calcite and dolomite cements that
phology of the different types of siderites can be found in Rosenau et al. occur in interlaminated siltstones and sandstones and (2) sphaerosiderite.
(2013). Here, we summarize the major characteristics of the different The occurrence and morphology of the micritic siderite cements indicate
types of siderites that were analyzed for this study. The micritic form of that it most likely formed as a diagenetic alteration product in an early
siderite most commonly occurs in thinly interlaminated calcareous burial environment (Mozley 1989; Mozley and Carothers 1992).
siltstones and sandstones underlying paleosols, where it occurs with Pedogenic sphaerosiderite is a common mineral found in poorly drained,
micritic to coarsely crystalline calcite cements. Micritic siderite often reducing soils (Ludvigson et al. 1998; Ufnar et al. 2002, 2004; White et al.
defines bifurcating tubular networks that disrupt sedimentary bedding, 2001), and its occurrence has been documented in modern soils (Brewer
and more rarely occurs as thin coatings on calcium carbonate nodules in 1964; Postma 1981, 1982; Driese et al. 2010) as well as paleosols (Leckie et
paleosol horizons. Sphaerosiderite is more commonly associated with al. 1989; Mozley and Carothers 1992; Ludvigson et al. 1998).
well-developed gleyed B horizons of paleosol profiles and more rarely is Sphaerosiderites typically exhibit a narrow intraprofile range of d18O
found in interlaminated calcareous siltstones and sandstones. Sphaer- values and have been interpreted in other studies to reflect the average
osiderites occur as radial crystals organized into discrete spherules up to d18O value of the local ground waters from which they crystallized under
0.5 mm in diameter and as intergrown aggregates, or clusters, of a relatively narrow range of temperatures (Ludvigson et al. 1998; White et
spherules. Sphaerosiderites are concentrically zoned, and display alter- al. 2001; Ufnar et al. 2002, 2004). Pedogenic sphaerosiderite d13C values
nating zones of siderite and calcite. They typically display small cores typically display a large intraprofile range of up to 30% that appears to
darkened by ferric oxides and larger anhedral crystals of siderite on be related to the large kinetic isotope fractionation between CO2 and CH4
exteriors. The cores of many sphaerosiderite crystals contain inclusions of generated by methane-reducing bacteria in water-saturated, anoxic soil
opaque material that is likely framboidal pyrite and/or pyritized organic profiles (Faure et al. 1995; Ludvigson et al. 1998). This process can
matter. produce d13C values of CO2 up to +7% (Whiticar et al. 1986). These
values may be produced under closed-system, one-component CO2
Isotope Geochemistry mixing, as well as additional contributions of CO2 from oxidation of
methanogenic CH4 (Irwin et al. 1977; Ludvigson et al. 1998; Sheldon and
Siderite Oxygen and Carbon Isotopes.—Sphaerosiderite d18OV-PDB Tabor 2009). As such, d13C values of siderites likely reflect local
values range from 23.6% to 20.5% while micritic siderite cement environmental conditions and are not useful as a proxy for regional
d18OV-PDB values range from 24.7% to 20.6% (Fig. 14A, Tables 2, 3). paleoclimate reconstruction. Rather, the discussion herein focuses on the
Sphaerosiderite d18O values display intraprofile variation where B- intraprofile and interprofile variation of sphaerosiderite d18O values and
horizon sphaerosiderites are consistently +0.5% to +1% more negative massive siderite d18O values in conjunction with the d18O and dD values
than sphaerosiderites in the underlying C or R horizons. Sphaerosiderite of a coeval uppermost Desmoinesian (Shelburn Formation) flint-clay
and massive siderite cement d13CV-PDB values range from 212.8% to kaolinite as a proxy for middle–upper Pennsylvanian paleoenvironmental
23.2% and 217.9% to +2.2%, respectively (Fig. 14B, Tables 2, 3). conditions.
Flint-Clay Kaolinite Oxygen and Hydrogen Isotopes.—The average Interprofile and Intraprofile Variations of Siderite d18O Values.—In
18
d OV-SMOW and dDV-SMOW values of kaolinite from the , 0.2 mm size general, d18O values of micritic siderite cement from interlaminated
fraction from the flint-clay breccia are +21.0% and 246.5%, respectively siltstones and sandstones stratigraphically underlying paleosol profiles
(Table 4). The wt % of the 100uC H2O (sorbed H2O) ranges from 2.2% to show no distinct stratigraphic distribution and display a wide range of
3.9%, while the 825uC H2O (structural H2O) ranges from 13.3% to 15.0%. d18O values that do not define any obvious temporal trends (Fig. 14A,
The wt % H2O in stoichiometric kaolinite [Al2Si2O5(OH)4 ] is 13.96%. As Table 3). These micritic siderite cements are interpreted to represent
such, the calculated wt % H2O of 15.0% from one of the analyses of the diagenetic alteration products that formed soon after burial of the
flint-clay kaolinite may indicate incomplete removal of sorbed H2O from sediments, prior to compaction and lithification. A number of possible
the sample or another mineral present that is characterized by a higher wt mechanisms during diagenesis could explain the variability observed in
% H2O, but in such a low concentration it is not detected by XRD micritic siderite d18O values, including recrystallization, water–mineral
analysis, such as gibbsite (Tabor and Yapp 2005). The wt % of O2 in the interactions, mixing of meteoric and marine waters, or precipitation at
kaolinite samples ranges from 49 to 54%, which is slightly lower than the very high temperatures (Mozley and Carothers 1992; Mozley and Burns
wt % O2 in stoichiometric kaolinite (55.8%). 1993). Regardless of the precise mechanism(s) responsible for their
r
FIG. 7.—Measured paleosol-bearing sections of the middle–upper Pennsylvanian Shelburn Formation. Measurements depicted on core graphic logs are in meters (bold
text) and feet (italicized text) as measured from the top of the core. Black and white bar scale is in increments of 10 feet. Wavy lines to the left of measured sections define
tops of paleosols. Capital letters refer to paleosols of a given pedotype. Encircled capital letters refer to pedotype discussed in Rosenau et al. (2013). Encircled numbers
and acronyms above measured sections represent site identification (Fig. 1). Positions of sphaerosiderite and micritic siderite samples are denoted by red and black stars,
respectively. Correlations of sections follow Willman et al. (1975) and Nelson et al. (2011).
650 N.A. ROSENAU ET AL. JSR
FIG. 10.—Measured paleosol-bearing sections of the upper Pennsylvanian Mattoon Formation. Measurements depicted on core graphic logs are in meters (bold text)
and feet (italicized text) as measured from the top of the core. Black and white bar scale is in increments of 10 feet. Wavy lines to the left of measured sections define tops
of paleosols. Capital letters refer to paleosols of a given pedotype. Encircled capital letters refer to pedotype discussed in Rosenau et al. (2013). Encircled numbers and
acronyms above measured sections represent site identification (Fig. 1). Positions of micritic siderite samples are denoted by black stars, respectively. Correlations of
sections follow Willman et al. (1975) and Nelson et al. (2011).
and Gilg 1996), resulting in a calculated meteoric water d18O value of above and below the D-M boundary, using the experimentally
23.1%. This temperature estimate is assumed to be a reasonable determined temperature-dependent fractionation factor for siderite
approximation of mean annual surface temperatures at the time of (Fig. 15; Carothers et al. 1988). Using the independently derived
kaolinite crystallization because of the large thermal-mass of soil paleotemperature estimate from the flint-clay kaolinite and assuming
compared to the atmosphere and because temperature in the low-latitude that coexisting sphaerosiderites crystallized near chemical equilibrium
tropics because temperatures in the low-latitude tropics often do not vary permits an additional estimate of late middle Pennsylvanian meteoric
by more than 1–2uC through an annual cycle (Barron and Moore 1994; water d18O values. When a crystallization temperature estimate of
Barron and Fawcett 1995; Buol et al. 1997). Soil–water d18OV-SMOW 27 6 2uC is plotted in conjunction with measured d18O values from
values were calculated over a range of temperatures (0–40uC) from the coeval sphaerosiderite, an estimated range of middle–late Pennsylvanian
average sphaerosiderite d18O values associated with the flint-clay meteoric-water d18O values from 22.7% to 23.6%, with an average
kaolinite and from stratigraphically bounding paleosols immediately value of 23.1%, is calculated equivalent to the meteoric-water d18O value
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 653
calculated from coeval flint-clay kaolinite d18O and dD values (Fig. 15; Protosols) paleosols with relatively weak morphological development and
Table 4). The range of meteoric-water d18O values calculated from the horizonation in the lower–middle Pennsylvanian strata followed by the
d18O values of paired flint-clay kaolinite and sphaerosiderite correspond diversity of paleosol types (Types C-G) with well-developed morphologies
to kaolinite crystallization temperatures ranging from 30uC to 24uC, and horizonation in the middle–upper Pennsylvanian is an indication of
respectively, and an average of 27uC. an overall increase in landscape stability and decreased sediment
accumulation rates during intervals of pedogenesis (Figs 4–10, 16, 18;
Autogenic Processes: The Role of Geomorphology on the Lateral and Kraus and Aslan 1993; McCarthy et al. 1997; Kraus and Aslan 1999;
Stratigraphic Distribution of Illinois Basin Paleosols McCarthy et al. 1999; Demko et al. 2004; Tabor et al. 2004). These
intervals of greater landscape stability occurred during eustatic low-
Lateral Trends.—Types C, D, and E paleosols (gleyed Vertisols, gleyed stands, which is concomitant with the onset and duration of high-order
calcic Vertisols, and gleyed vertic Calcisols, respectively) exhibit distinct transgressive–regressive cycles and the deposition of classical Pennsylva-
lateral variations in morphology from the northern margin to the interior nian cyclothemic strata. The lower Pennsylvanian stratigraphy of the IB
of the IB (Figs. 4–10, 16). Specifically, the observation of high-chroma is principally composed of fluvial sandstones, and minor amounts of
hues of paleosol matrix from the basin margin (LSC core) in conjunction marine and terrestrial siltstones and shales, and marine limestones
with gleyed, siderite- and pyrite-rich, age-equivalent deposits from (Hopkins and Simon 1975; Siever 1957), as well as many upward-fining
sequences. Lower Pennsylvanian clastic input into the IB appears to have
paleosols in the basin interior (MAC, ELY, SAL, FWL cores and field
been depositionally influenced by the pre-existing Mississippian–Penn-
sites) indicate that soil development was strongly influenced by
sylvanian topographic unconformity as evidenced by the geographic
differences in soil hydrology that are similar to a modern soil catena,
distribution of large sandstone bodies in lower Pennsylvanian strata. The
whereby the paleosols represent soil development upon similar parent
preservation of these large sandstone bodies in conjunction with thin and
material under similar climate regimes, and the different morphological
poorly developed coal seams and common compound Type B paleosols
and mineralogical characteristics between laterally continuous, age-
indicate that those paleolandscapes were subject to relatively high rates of
equivalent profiles are due to variations in relief and drainage (Figs. 16,
sediment accumulation that disrupted soil formation and led to isolated
17; Jenny 1980; Birkeland 1984; Besly and Fielding 1989; Miller and
swamplands of peat accumulation. Nevertheless, the presence of gleyed
Birkeland 1992; Wright and Marriott 1993; McCarthy et al. 1997;
matrix colors and coal indicate that they were likely forming under
McCarthy and Plint 2003; Tabor et al. 2006). As such, autogenic poorly-drained conditions in a humid, ever-wet environment (Cecil et al.
processes that resulted in paleocatenary relationships may best explain 1985, 2003; Ziegler et al. 1987; Lézine and Chateauneuf 1991). As such,
basin-scale trends in Types C, D, and E paleosols (Figs. 16, 17). we interpret the stratigraphic distribution of paleosol types of the lower
Lateral variations in paleosol morphology and redox-sensitive mineral Pennsylvanian in the IB to result principally from changes in landscape
assemblages of age-equivalent Types C, D, and E paleosols from stability and sediment accumulation rates.
different parts of the IB record are interpreted to result from A shift in the stratigraphic distribution of paleosol types occurs in the
dominantly local (basin-scale) processes such as differences in topog- middle Pennsylvanian Carbondale Formation. At this level, and above,
raphy and from relative rates of subsidence between these two portions Types C (gleyed Vertisols), D (gleyed calcic Vertisols), E (gleyed vertic
of the basin. These features led to overall better-drained conditions Calcisols), F (gleyed Calcisols), and G (vertic Calcisols) paleosols become
throughout the developmental history of the paleosols in the northern an important stratigraphic component of the IB (Figs. 4–10, 16, 18). This
part of the IB (LSC core) and a shallower water table and more change in paleosol distribution is significant because it indicates an overall
frequent marine flooding in the interior of the basin (Fig. 17). increase in landscape stability and soil drainage and a decrease in sediment
Furthermore, the restricted occurrence of Type G (calcic Vertisols) accumulation rates during some parts of cyclothem development.
paleosols along the northern margin (Figs. 7–9) and the observation
that age-equivalent profiles from the basin interior preserve evidence for Allogenic Processes: The Role of Glacioeustasy and Associated Climate
multi-stage (oxygenated and reduced) polygenesis (Rosenau et al. 2013) Change on Paleosol Development
is interpreted to be the result of basin-scale topographic differences
across the basin. Specifically, it suggests that lower landscapes in the Tectonic Models.—Tectonic models have been put forth to explain the
basin interior were more prone to gley processes as a result of ground- formation of Pennsylvanian cyclothems in the IB (Klein and Willard
water, brackish-water, and sea-water inundation (Fig. 17). Importantly, 1989) and other North American basins (Quinlan and Beaumont 1984;
the patterns of calcite, siderite, and pyrite cementation as well as illite– Tankard 1986; Klein and Cloetingh 1989). Tectonic highlands likely
smectite ordering among these paleosol profiles (Rosenau et al. 2013) influenced local climate patterns, and tectonic subsidence may have
support this landscape model. resulted in minor relative sea-level changes and provided accommodation
space for cyclic sedimentation in the IB and surrounding basins (Weller
Stratigraphic Trends.—The stratigraphic distribution of paleosols 1931; Klein and Kupperman 1992). However, the numerous, widespread,
similarly defines a trend that can be related to basin-scale autogenic and repetitious marine and nonmarine cycles across Euramerica, the
processes. The occurrence of only Type A (Histosols) and B (gleyed short interval between many of these cycles, and the well-established
r
FIG. 12.—Photographs and photomicrographs of the flint-clay breccia layer (FCB; Fig. 3, FCB at 79.8–82.1 m. A) Photograph of MAC core at , 82.1 m showing
siderite (s) and flint-clay (fc) in the FCB. B) Photograph (close up) of MAC core at , 82.4 m showing siderite (s) and flint-clay (fc) in FCB. C) Photograph of cross-
section of core from the FCB interval showing flint-clay (fc) matrix and sphaerosiderite (s). D) Photomicrograph (cross-polarized light) showing euhederal dolomite (d)
crystals with some ferric oxide staining in flint-clay matrix (fc). E) Photomicrograph (plane-polarized light) showing euhderal dolomite (d) crystals with ferric-oxide-
stained cores in flint-clay matrix (fc). Note that dolomite disrupts internal bedding of flint-clay. F) Photomicrograph (cross-polarized light) illustrating association of
siderite (s), micrite (m), and sparry calcite (sp) in the FCB. Note that sparry calcite and siderite crosscut micrite, indicating that they formed later.
656 N.A. ROSENAU ET AL. JSR
FIG. 14.—Stratigraphic distribution of the oxygen and carbon isotope composition of sphaerosiderite and micritic siderite cements. Filled circles represent samples
from morphologically well-developed pedogenic B horizons in paleosol profiles. Open circles represent samples from C or R horizons of paleosol profiles. Open triangles
represent samples taken from the flint-clay breccia (FCB) layer in the MAC core (Fig. 7). Rectangular boxes encompass sets of samples from a single paleosol profile. A)
Stratigraphic distribution of siderite d18OV-PDB values. B) Stratigraphic distribution of siderite d13CV-PDB values. Note consistently more negative d18O values of B-
horizon sphaerosiderites with respect to underlying C-horizon sphaerosiderites. Generalized stratigraphic column modified from Willman et al. (1975) and Nelson et
al. (2011).
redox reactions that promoted precipitation of pyrite, siderite, and gley amount and frequency of low-latitude precipitation in conjunction with
matrix colors (Fig. 19B). Depending on the availability of sediment and glacio-eustatic changes in sea level. In particular, following the model of
local topography, a laminated organic-rich, kaolinitic layer was Cecil et al. (2003) it appears that IB soils began actively forming under
deposited. Continued groundwater rise due to increased precipitation seasonal precipitation during early glacial intervals (late regression).
eventually led to peat-swamp development on top of the paleosol and in- Subsequently during glacial maximum (lowstand), when precipitation
situ accumulations of plant organic matter (Fig. 19B). During glacial increased and climate became less seasonal, soils became water-saturated
termination, sea level and base level rose. Types B, C, D, and E paleosols and anoxic, leading to precipitation of reduced-iron minerals (Fig. 19).
were flooded by brackish, and eventually marine, water and ultimately Morphological and mineralogical evidence from IB paleosols suggests
buried by marine strata. Consequently, we consider most individual that changes in low-latitude precipitation and variations in sea level were
paleosol profiles in the IB to document a complex paleohydrologic record closely linked. Based on the evidence presented herein, we propose that
that led to changes in the degree of hydromophy, driven by changes in the while ultimately both sea-level rise and increases in the amount and
658 N.A. ROSENAU ET AL. JSR
TABLE 2.— d18OV-PDB and d13CV-PDB values of sphaerosiderite associated with paleosols and flint-clay breccia.
frequency of precipitation were responsible for the final preservation of American basins during the later middle and late Pennsylvanian
redox-sensitive paleosol features, the initial rise in the paleo–water table (DiMichele et al. 2010) a similar climate-driven model may be applicable
and peat-swamp development, was due to an increase in the amount and to other paleosol-bearing basins of west central (Appalachian foreland
frequency of precipitation over the low-gradient North American basin) and west equatorial (Midcontinent basin) Pangaea.
landscape as a result of confinement and intensification of the ITCZ
near the equator (Cecil et al. 2003; Poulsen et al. 2007; Elrick and Nelson Middle–late Pennsylvanian (Desmoinesian–Missourian) Climatic and
2010). In other words, relative rises in sea level did not force concomitant Glacioeustatic Changes: Effects on the Stratigraphic Distribution of
rises in the water table, but instead, rises in the water table preceded, and
Paleosol Types
outpaced, sea-level rise in the IB, reaching a maximum just before, and
during peat accumulation, in response to increases in the amount and The middle–late Pennsylvanian (, Desmoinesian–Missourian, D–M)
intensity of precipitation (Fig. 19). This is consistent with interpretations boundary is a significant interval in the Pennsylvanian marked by abrupt
regarding the formation of thick, widespread, laterally traceable coal beds evolutionary change in large numbers of marine and terrestrial species.
in middle and upper Pennsylvanian strata among North American basins Studies of marine taxa (i.e., fusulinids, brachiopods, ammonoids, and
(Cecil et al. 2003; Greb et al. 2003; Elrick and Nelson 2010). Furthermore, conodonts; Heckel and Weibel 1991; Rosscoe 2008), terrestrial plants
due to the extremely flat topography and connectedness of the North (DiMichele and Phillips 1996), and tetrapods (Berman et al. 2010; Sahney
TABLE 3.— d18OV-PDB and d13CV-PDB values of micritic siderite associated with interlaminated siltstones and sandstones.
TABLE 4.— dDV-SMOW and d18OV-SMOW values of the latest Desmoinesian flint-clay.
FIG. 16.—Schematic stratigraphic sections of MAC, LSC, and ELY cores highlighting the stratigraphic positions and lateral distribution of different paleosol types
(filled rectangles). Lateral changes in paleosol types among age-equivalent deposits are interpreted to represent morphological changes associated with a paleocatena
(Rosenau et al. 2013). Different colors represent different paleosol types, as in Figure 4. Note different scale for ELY core. Correlations of sections follow Willman et al.
(1975) and Nelson et al. (2011).
periods of modest-amplitude incision events, indicating eustatic base-level 2008c) as well as far-field records in western North America (Bishop et al.
changes ranging from . 60 m to , 30 m (Rygel et al. 2008; Fischbein et 2010) and high-latitude western Gondwanaland (Gulbranson et al. 2010)
al. 2009). suggests that aridification across the D–M boundary was due not to an
Accordingly, an alternative position on D–M eustatic change that is intense glacial phase, but instead to greenhouse warming and attendant
based on near-field Gondwanan records (Fielding et al. 2008a, 2008b, waning of high-latitude Gondwanan ice sheets. This interpretation is
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 661
generally consistent with climate models, which suggest that low-latitude first occurrence of a Type F paleosol is below the Chapel Coal (Figs. 8,
aridification was caused by weaker Hadley cell convection and southward 16). Type F paleosols lack shrink–swell features, and they exhibit a clay
drift of the ITCZ during periods of minimum glacial ice (Poulsen et al. mineralogy that noticeably lacks any relict 2:1 expandable phyllosilicate
2007). Regardless of the precise mechanism(s) for low-latitude climate component; this may be the result of a shift towards overall drier
change there is a growing consensus that tropical paleoclimate underwent conditions. These features are in stark contrast to those that commonly
a transition to drier, warmer, and more seasonal conditions (Cecil et al. characterize other well-developed paleosols in the IB. The occurrence of
2003; Pfefferkorn et al. 2008) across the middle–late Pennsylvanian Type F paleosols is restricted to only three examples in the lower
boundary perhaps on the timescale of one cyclothem (, 104 to 105 years; Missourian (Patoka and Bond Formations) and is found in only one core
Heckel 2008). (ELY core) from the basin interior (Figs. 1, 8, 16). Furthermore, age-
Similarly, the stratigraphic distribution of Types F and G paleosols in equivalent soils from the LSC and MAC core display shrink–swell
the IB exhibit characteristic features that indicate a significant shift in the features and calcium carbonate accumulation, indicating that they
magnitude and duration of seasonality (Figs. 16–18). Paleosols indicative formed under some degree of seasonal precipitation (Figs 16, 18; Yaalon
of formation under seasonal precipitation are found before, and after, the and Kalmar 1978; Wilding and Tessier 1988; Buol et al. 1997).
D–M transition and show no distinct changes in morphology or The first occurrence of Type G paleosols occurs in lower Missourian
distribution across the boundary, suggesting continued seasonal rainfall strata (Shelburn Formation) below the Danville Coal (Figs. 16, 18).
patterns in the late Pennsylvanian. However, Types F and G paleosols Examples of this paleosol type come from the basin-margin LSC core and
appear in the IB in uppermost lower Missourian strata (Figs. 16, 18). The one occurrence in the MAC core (Figs. 7–9, 18). As discussed previously,
662 N.A. ROSENAU ET AL. JSR
FIG. 18.—Diagrams illustrating the distribution of paleosol types through time (middle to late Pennsylvanian) and across the Illinois basin (basin margin to interior).
Different colors represent different paleosol types, as in Figure 4. Encircled numbers and acronyms above measured sections represent site identification (Fig. 1).
these paleosols are generally thicker and are characterized by relict- on contemporaneous paleo-floral assemblages (Phillips 1979; Blake et al.
chroma red hues, well-developed shrink–swell features, significant 1999; Pfefferkorn et al. 2008; Falcon-Lang and DiMichele 2010; Sahney
accumulations of calcium carbonate, and lack of evidence for long et al. 2010) and geological evidence (Cecil et al. 1985; Cecil 1990; Cecil et
intervals of waterlogging like age-equivalent paleosols from the basin al. 2003; Olszewski and Patzkowsky 2003; Feldman et al. 2005; Bishop et
interior. Furthermore, the , 2 mm clay-mineral assemblage of IB al. 2010; Gulbranson et al. 2010). As such, the influence of low-latitude
paleosols exhibits distinct stratigraphic trends in the relative abundance regional climate, ultimately driven by high-latitude ice-sheet dynamics
of kaolinite, I/S, and illite in paleosols from the MAC and LSC cores. and changes in the position and size of the ITCZ, provides a reasonable
Specifically, there is a significant decrease in the relative abundance of explanation for the observed changes in the stratigraphic distribution of
kaolinite in the , 2 mm fraction in paleosols from the MAC and LSC paleosol types across the D–M boundary.
cores near the D-M boundary (Rosenau et al. 2013).
The morphology and relative clay-mineral abundances among paleo- Siderite and Kaolinite as Paleoenvironmental Proxies across the
sols (Rosenau et al. 2013) in the IB define a stratigraphic trend that is
Desmoinesian-Missourian Boundary
interpreted to reflect a period of overall decreased hydromorphy,
increased free-drainage and landscape stability, and a shift in the overall Considering that sphaerosiderite d18O values were obtained from
regional climate to drier conditions in the Upper Pennsylvanian successive cyclothems across the D-M boundary it is possible to discuss,
(Missourian), which are consistent with climate interpretations based at the timescale of individual cyclothems, the mechanisms that were
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 663
FIG. 19.—Model illustrating the development of lithostratigraphic components of an idealized Pennsylvanian cyclothem in the Illinois basin with emphasis on factors
affecting the development of polygenetic paleosols from the interior of the basin. A) Diagram depicting the development of major stratigraphic components of an
idealized Pennsylvanian cyclothem from the interior of the Illinois basin with respect to glacial–interglacial changes in sea level (solid black line) and water-table position
(gray dashed line) in conjunction with sequence stratigraphic interpretations. Abbreviations for systems tracts: HST, highstand systems tract; RST, regressive systems
tract; LST, lowstand systems tract; TST, transgressive systems tract. The rise in the water table is interpreted to occur during the late stage of the RST, prior to the rise in
sea level associated with the TST. The rise in the water-table is interpreted to begin during the end, or soon after, the cessation of pedogenesis, reaching a maximum
during peat formation. B) Proposed three-stage (I, II, III) model for the development of polygenetic paleosol profiles in the interior of the Illinois basin (ELY, MAC,
FWL, and SAL cores, and field sites) within the context of glacial–interglacial sea-level changes and water-table position as depicted in Part A.
potentially responsible for the observed shift in sphaerosiderite d18O terms of either changing d18O meteoric-water values or temperature
values across the D-M boundary. As discussed previously, trends defined (Fig. 15). If the , 1.5% negative shift in sphaerosiderite d18O values is to
by the D-M sphaerosiderite d18O values in conjunction with a latest be explained solely by a temperature change, this would suggest a
Desmoinesian kaolinite paleotemperature estimate can be explained in temperature change of approximately + 6uC across the D-M boundary,
664 N.A. ROSENAU ET AL. JSR
occurring over the time scale of one cyclothem, suggesting temperatures sphaerosiderite d18O values occurring on the order of one cyclothem
in the soil ecosystems of the IB that were 33 6 2u C in the early spanning the Desmoinesian–Missourian boundary followed by a return
Missourian (Figs. 14, 15). Similarly, the shift back to more positive values to more positive siderite d18O values, also on the order of one cyclothem,
in the early Missourian suggests a cooling of 6uC, also occurring over the is interpreted to represent a relatively short-lived temperature increase of
time scale of one cyclothem (Figs. 14, 15). Alternatively, if temperature is , 6uC across the Desmoinesian–Missourian boundary. Collectively,
assumed to have remained relatively constant across the D-M interval, these proxy records reveal significant paleoenvironmental changes across
the shift in sphaerosiderite d18O values necessitates a mechanism that the middle–late Pennsylvanian (Desmoinesian–Missourian) in the tropics
could cause soil-water d18O values to become depleted by 1.5% over the of North America which include (1) an increase in free-drainage and
timescale of a single cyclothem. Such a mechanism could be a transient landscape stability, (2) a shift in the overall regional climate to drier
period of increased precipitation, leading to 18O-depleted air masses and conditions, and (3) a transient period of elevated low-latitude tempera-
the negative sphaerosiderite d18O values associated with the D-M tures in the early late Pennsylvanian.
boundary strata. Given these possible scenarios, a short-lived tempera-
ture increase of # 6uC across the D-M boundary is the preferred ACKNOWLEDGMENTS
explanation for the siderite d18O values reported here. This interpretation
is consistent with previous studies that have postulated the D-M The authors would like to thank Phil Ames of Peabody Energy, Barry
Sargeant of Knight Hawk Coal, Vigo Coal Operating Company, and Triad
boundary interval represents an overall warming of Earth’s climate and Mining, Inc. for providing access to their mines and permitting collection of
drier conditions in the paleotropics (Frakes et al. 1992; Cleal and Thomas samples, as well as the Illinois State Geological Survey for providing access to
2005; Pfefferkorn et al. 2008). A similar process may be responsible for cores, core descriptions, and core samples. We thank Drs. Paul McCarthy,
the observed sphaerosiderite d18O values in the early Missourian. Celina Suarez, Fred Longstaffe, and Julia Kahmann-Robinson for helpful
Specifically, the return to more positive sphaerosiderite d18O values in reviews that considerably improved the quality of the manuscript. N.
the early Missourian could be indicative of soil waters that have Rosenau was supported by a National Science Foundation Graduate
Research fellowship, a Roy M. Huffington fellowship provided by the Roy
undergone evaporative enrichment (Cerling 1984; Cerling and Quade M. Huffington Department of Earth Sciences (SMU), as well as grants from
1993; Rozanski et al. 1993; Tabor et al. 2002) due to increased aridity in the Geological Society of America and The Clay Minerals Society. N. Tabor
the low-latitude tropics of Pangaea. Unfortunately, well-preserved was supported by NSF-EAR 0545654 and NSF-EAR 0844147.
pedogenic sphaerosiderites are absent in the studied stratigraphic sections
above this interval, and as such it is not possible to determine whether this REFERENCES
trend continues up section.
AITKEN, J.F., AND FLINT, S.S., 1996, Variable expressions of interfluvial sequence
boundaries in the Breathitt Group (Pennsylvanian), eastern Kentucky, USA, in
CONCLUSIONS Howell, J.A., and Aitken, J.F., eds., High Resolution Sequence Stratigraphy:
Innovations and Applications: Geological Society of London, Special Publication
Middle–upper Pennsylvanian strata in the Illinois basin exhibit 104, p. 193–206.
prominent lateral and stratigraphic trends in paleosol morphology and ALLGAIER, G.J., AND HOPKINS, M.E., 1975, Reserves of the Herrin (No. 6) Coal in the
clay mineralogy that reflect the influence of local to global-scale Fairfield Basin in southeastern Illinois: Illinois Geological Survey, Circular 489, 31 p.
ANDRESEN, M.J., 1961, Geology and petrology of the Trivoli Sandstone in the Illinois
processes. The preservation of high-chroma hues of paleosols from the Basin: Illinois State Geological Survey, Circular 316, 31 p.
basin margin in conjunction with gleyed, siderite- and pyrite-rich age- ARCHER, A.W., 1998, Hierarchy of controls on cyclic rhythmite deposition: Carbon-
equivalent deposits from paleosols in the interior of the basin indicate iferous basins of eastern and mid-continental USA, in Alexander, C.R., Davis, R.A.,
and Henry, V.J., eds., Tidalites: Processes and Products: SEPM, Society for
that soil development was strongly influenced by differences in soil Sedimentary Geology, Special Publication 61, p. 59–68.
hydrology as a result of autogenic factors, including variations in ARCHER, A.W., KUECHER, G.J., AND KVALE, E.P., 1995, The role of tidal-velocity
topography and relative rates of subsidence between the northern and asymmetries in the deposition of silty tidal rhythmites (Carboniferous, Eastern
Interior Coal Basin, U.S.A.): Journal of Sedimentary Research, v. 65, p. 408–416.
interior portions of the basin. Pedogenic features in paleosols from the BARRON, E.J., AND FAWCETT, P.J., 1995, The climate of Pangaea: a review of climate
basin interior indicate initial formation in well-drained environments model simulations of the Permian, in Scholle, P.A., Peryt, T.M., and Ulmer-Scholle,
under seasonal precipitation and a subsequent period of waterlogging and D.S., eds., The Permian of Northern Pangaea: Berlin, Springer, v. 1, p. 37–52.
BARRON, E.J., AND MOORE, G.T., 1994, Climate Model Application in Palaeoenviron-
development of reducing conditions throughout the soil profile. These mental Analysis: SEPM, Short Course 33, 339 p.
features are indicative of a multistage, polygenetic pedogenic history, as a BAUER, R.A., AND DEMARIS, P.J., 1982, Geologic investigation of roof and floor strata:
result of formation under contrasting chemical environments that are Longwall demonstration, Old Ben Mine No. 24: Illinois State Geological Survey,
Contract–Grant Report 1982-2, 49 p.
interpreted to have been driven by changes in Gondwanan ice volume and BERMAN, D.S., HENRICI, A.C., BREZINSKI, D.K., AND KOLLAR, A.D., 2010, A new
associated changes in the amount and frequency of precipitation due to trematopid (Temnospondyl; Dissorophoidea) from the Upper Pennsylvanian of
the equatorial confinement or intensification of the Intertropical western Pennsylvania: earliest evidence of terrestrial vertebrates responding to a
warmer, drier climate: Carnegie Museum Annals, v. 78, p. 289–318.
Convergence Zone. A change in the stratigraphic distribution of paleosol BESLY, B.M., AND FIELDING, C.R., 1989, Palaeosols in Westphalian coal-bearing and
types and the relative abundance of kaolinite in the clay fraction of red-bed sequences, central and northern England: Palaeogeography, Palaeoclimatol-
Illinois basin paleosols occurs near the Desmoinesian–Missourian ogy, Palaeoecology, v. 70, p. 303–330.
BESTLAND, E.A., RETALLACK, G.J., AND SWISHER, III, C.C., 1997, Stepwise climate
boundary. Specifically, paleosols preserving relict-chroma red hues, change recorded in Eocene–Oligocene paleosol sequences from central Oregon:
well-developed shrink–swell features, and significant accumulation of Journal of Geology, v. 105, p. 153–172.
calcium carbonate, and lacking evidence for long intervals of waterlog- BIRD, M.I., AND CHIVAS, A.R., 1988, Stable isotope evidence for low-temperature
ging, first occur in the lower Missourian. This is coincident with a weathering and post-formation hydrogen-isotope exchange in Permian kaolinites:
Chemical Geology, Isotope Geoscience Section, v. 72, p. 249–265.
significant decrease in the relative abundance of kaolinite in the , 2 mm BIRD, M.I., AND CHIVAS, A.R., 1989, Stable isotope geochronology of the Australian
fraction in paleosols from the MAC and LSC cores near the regolith: Geochimica et Cosmochimica Acta, v. 53, p. 3239–3256.
Desmoinesian–Missourian boundary indicative of a trend towards overall BIRKELAND, P.W., 1984, Soils and Geomorphology, Third Edition: New York, Oxford
University Press, 430 p.
more arid conditions. BISHOP, J.W., MONTAÑEZ, I.P., GULBRANSON, E.L., AND BRENCKLE, P.L., 2009, The onset
The d18O values of paleosol B-horizon sphaerosiderites in conjunction of mid-Carboniferous glacioeustasy: sedimentologic and diagenetic constraints,
with d18O and dD values of a latest Desmoinesian flint-clay kaolinite Arrow Canyon, Nevada: Palaeogeography, Palaeoclimatology, Palaeoecology, v.
276, p. 217–243.
provide additional evidence for a perturbation of the climate system BISHOP, J.W., MONTAÑEZ, I.P., AND OSLEGER, D.A., 2010, Dynamic Carboniferous
across the Desmoinesian–Missourian boundary. A negative 1.5% shift in climate change, Arrow Canyon, Nevada: Geosphere, v. 6, p. 1–34.
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 665
BISNETT, A.J., AND HECKEL, P.H., 1996, Sequence stratigraphy helps to distinguish DRIESE, S.G., LUDVIGSON, G.A., ROBERTS, J.A., FOWLE, D.A., GONZALEZ, L.A., SMITH,
offshore from nearshore black shales in the Midcontinent Pennsylvanian succession, J.-.J, VULAVA, V.M., AND MCKAY, L.D., 2010, Micromorphology and stable-
in Witzke, B.J., Ludvigson, G.A., and Day, J., eds., Paleozoic Sequence Stratigraphy: isotope geochemistry of historical pedogenic siderite formed in PAH-contaminated
Views from the North American Craton: Geological Society of America, Special alluvial clay soils, Tennessee, U.S.A.: Journal of Sedimentary Research, v. 80,
Paper 306, p. 341–350. p. 943–954.
BLAKE, B.M., JR., CROSS, A.T., EBLE, C.F., GILLESPIE, W.H., AND PFEFFERKORN, H.W., ELLIOTT, W.C., SAVIN, S.M., DONG, H., AND PEACOR, D.R., 1997, A paleoclimate
1999, Selected plant megafossils from the Carboniferous of the Appalachian Region, interpretation derived from pedogenic clay minerals from the Piedmont Province,
Eastern United States: geographic and stratigraphic distribution, in Hillis, L.V., Virginia: Chemical Geology, v. 142, p. 201–211.
Henderson, C.M., and Bamber, E.W., eds., Carboniferous and Permian of the World: ELRICK, S.D., AND NELSON, W.J., 2010, Facies relationships of the Middle Pennsylvanian
Canadian Society of Petroleum Geologists, Memoir 19, p. 259–335. Springfield Coal and Dykersberg Shale: constraints on sedimentation, development of
BREWER, R.C., 1964, Fabric and Mineral Analysis of Soils: New York, Wiley, 470 p. coal splits and climate change during transgression [abstract]: Geological Society of
BROUGH, J., 1929, On rhythmic deposition in the Yoredale Series: University of America, Abstracts with Programs, v. 42, p. 51.
Durham, Philosophical Society, Proceedings, v. 8, Part 2, p. 116–126. FALCON-LANG, 2004, Pennsylvanian tropical forests responded to glacial–interglacial
BUOL, S.W., HOLE, F.D., MCCRACKEN, R.J., AND SOUTHARD, R.J., 1997, Soil Genesis and rhythms: Geology, v. 32, p. 689–692.
Classification, Fourth Edition: Ames, Iowa, Iowa State University Press, 527 p. FALCON-LANG, H.J., AND DIMICHELE, W.A., 2010, What happened to the coal forests
CAROTHERS, W.W., ADAMI, L.H., AND ROSENBAUER, R.J., 1988, Experimental oxygen during Pennsylvanian glacial phases?: Palaios, v. 25, p. 611–617.
isotope fractionation between siderite-water and phosphoric acid liberated CO2- FALCON-LANG, H.J., HECKEL, P.H., DIMICHELE, W.A., BLAKE, B.M, JR., EASTERDAY,
siderite. Geochimica et Cosmochimica Acta, v. 52, p. 2445–2450. C.R., EBLE, C.F., ELRICK, S., GASTALDO, R.A., GREB, S.F., MARTINO, R.L., NELSON,
CECIL, C.B., 1990, Paleoclimate controls on stratigraphic repetition of chemical and W.J., PFEFFERKORN, H.W., PHILLIPS, T.L., AND ROSSCOE, S.J., 2011, No major
siliciclastic rocks: Geology, v. 18, p. 533–536. stratigraphic gap exists near the middle–upper Pennsylvanian (Desmoinesian–
CECIL, C.B., STANTON, R.W., NEUZIL, S.G., DULONG, F.T., RUPPERT, L.F., AND PIERCE, Missourian) boundary in North America: Palaios, v. 26, p. 125–139.
B.J., 1985, Paleoclimate controls on late Paleozoic sedimentation and peat formation FAURE, K., HARRIS, C., AND WILLIS, J.P., 1995, A profound meteoric water influence on
in the central Appalachian Basin: International Journal of Coal Geology, v. 5, p. 195– genesis in the Permian Waterberg coalfield, South Africa: evidence from stable
230. isotopes: Journal of Sedimentary Research, v. 65, p. 605–613.
CECIL, C.B., EBLE, C.F., FEDORKO, N., BLAKE, B.M., AND GRADY, W.C., 1992, FELDMAN, H.R., FRANSEEN, E.K., JOECKEL, R.M., AND HECKEL, P.H., 2005, Impact of
Paleoclimate controls on Carboniferous sedimentation and cyclic stratigraphy in the longer–term modest climate shifts on architecture of high-frequency sequences
Appalachian Basin: United States Geological Survey, Open File Report 92-546. (cyclothems), Pennsylvanian of Midcontinent USA: Journal of Sedimentary Research,
CECIL, C.B., DULONG, F.T., COBB, J.C., AND SUPARDI, C., 1993, Allogenic and autogenic v. 75, p. 350–368.
controls on sedimentation in the central Sumatra basin as an analogue for FERM, J.C., 1970, Allegheny delta deposits, in Morgan, J.P., ed. Deltaic Sedimentation,
Pennsylvanian coal-bearing strata in the Appalachian Basin, in Cobb, J.C., and Modern and Ancient: SEPM, Special Publication 15, p. 246–255.
Cecil, C.B., eds., Modern and Ancient Coal-Forming Environments: Geological FERM, J.C., 1974, Carboniferous environmental models in eastern United States and
Society of America, Special Paper 286, p. 3–22. their significance, in Briggs, G., ed., Carboniferous of the southeastern United States:
CECIL, C.B., DULONG, F.T., WEST, R.R., STAMM, R., WARDLAW, B., AND EDGAR, N.T., Geological Society of America, Special Paper 148, p. 79–95.
2003, Climate controls on the stratigraphy of a middle Pennsylvanian cyclothem in FIELDING, C.R., FRANK, T.D., BIRGENHEIER, L.P., RYGEL, M.C., JONES, A.T., AND
North America, in Cecil, C.B., and Edgar, N.T., eds., Climate Controls on ROBERTS, J., 2008a, Stratigraphic imprint of the late Palaeozoic ice age in eastern
Stratigraphy: SEPM, Special Publication 77, p. 151–180. Australia: a record of alternating glacial and non-glacial climate regime: Geological
CERLING, T.E., 1984, The stable isotopic composition of modern soil carbonate and its Society of London, Journal, v. 165, p. 129–140.
relationship to climate: Earth and Planetary Science Letters, v. 71, p. 229–240. FIELDING, C.R., FRANK, T.D., BIRGENHEIER, L.P., RYGEL, M.C., JONES, A.T., AND
CERLING, T.E., AND QUADE, J., 1993, Stable carbon and oxygen isotopes in soil ROBERTS, J., 2008b, Stratigraphic record and facies associations of the late Paleozoic
carbonates, in McKenzie, J.A., and Savin, S., eds., Climate Change in Continental ice age in eastern Australia (New South Wales and Queensland), in Fielding, C.R.,
Isotopic Records: American Geophysical Union, Geophysical Monograph 78, p. 217– Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and
231. Space: Geological Society of America, Special Paper 441, p. 41–58.
CLAYTON, R.N., AND MAYEDA, T.K., 1963, The use of bromine pentafluoride in the FIELDING, C.R., FRANK, T.D., AND ISBELL, J.L., 2008c, The late Paleozoic ice age—
extraction of oxygen from iron oxides and silicates for isotopic analysis: Geochimica a review of current understanding and synthesis of global climate patterns,
et Cosmochimica Acta, v. 27, p. 43–52. in Fielding, C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic
CLEAL, C.J., AND THOMAS, B.A., 2005, Palaeozoic tropical rainforests and their effect on Ice Age in Time and Space: Geological Society of America, Special Paper 441,
global climates: Is the past the key to the present?: Geobiology, v. 3, p. 13–31. p. 343–443.
CRAIG, H., 1961, Isotopic variations in meteoric waters: Science, v. 133, p. 1702–1708. FISCHBEIN, S.A., JOECKEL, R.M., AND FIELDING, C.R., 2009, Fluvial–estuarine
DEMKO, T.M., CURRIE, B.S., AND NICOLL, K.A., 2004, Regional paleoclimatic and reinterpretation of large, isolated sandstone bodies in epicontinental cyclothems,
stratigraphic implications of paleosols and fluvial/overbank architecture in the Upper Pennsylvanian, northern Midcontinent, USA, and their significance for
Morrison Formation (Upper Jurassic), Western Interior, USA: Sedimentary Geology, understanding late Paleozoic sea-level fluctuations: Sedimentary Geology, v. 216, p.
v. 167, p. 115–135. 15–28.
DEVERA, J.A., 1989, Ichnofossil assemblages and associated lithofacies of the Lower FOLKOFF, M.E., AND MEENTEMEYER, V., 1987, Climatic control of the geography of clay
Pennsylvanian (Caseyville and Tradewater Formations), southern Illinois: Kentucky, minerals genesis: Association of American Geographers Annals, v. 77, p. 635–650.
Indiana, and Illinois Geological Surveys, Illinois Basin Studies 1, p. 57–83. FOOSE, R.M., 1944, High alumina clays of Pennsylvania: Economic Geology, v. 39, p.
DIMICHELE W.A., AND NELSON, W.J., 1989, Small-scale spatial heterogeneity in 557–577.
Pennsylvanian-age vegetation from the roof-shale of the Springfield Coal: Palaios, FRAKES, L.A., FRANCIS, J.E., AND SYKTUS, J.I., 1992, Climate Modes of the Phanerozoic:
v. 4, p. 276–280. Cambridge, U.K., Cambridge University Press, 274 p.
DIMICHELE, W.A., AND PHILLIPS, T.L., 1996, Climate change, plant extinctions, and GIBBS, M.T., REES, P.M., KUTZBACH, J.E., ZIEGLER, A.M., BEHLING, P.J., AND ROWLEY,
vegetational recovery during the middle–late Pennsylvanian transition: the case of D.B., 2002, Simulations of Permian climate and comparisons with climate-sensitive
tropical peat-forming environments in North America, in Hart, M.L., ed., Biotic sediments: Journal of Geology, v. 110, p. 33–55.
Recovery from Mass Extinction Events: Geological Society of America, Special GONFIANTINI, R., 1984, Advisory group meeting on stable isotope reference samples for
Publication 102, p. 201–221. geochemical and hydrological investigations: Report, International Atomic Energy
DIMICHELE, W.A., EBLE, C.F., AND CHANEY, D.S., 1996, A drowned lycopsid forest Agency, Vienna.
above the Mahoning coal (Conemaugh Group, Upper Pennsylvanian) in eastern GREB, S.F., ANDREWS, W.M., EBLE, C.F., DIMICHELE, W., CECIL, C.B., AND HOWER,
Ohio, U.S.A.: International Journal of Coal Geology, v. 31, p. 249–276. J.C., 2003, Desmoinesian coal beds of the Eastern Interior and surrounding basins:
DIMICHELE, W.A., TABOR, N.J., CHANEY, D.S., AND NELSON, W.J., 2006, From wetlands the largest tropical peat mires in earth history, in Chan, M.A., and Archer, A.W., eds.,
to wet spots: environmental tracking and the fate of Carboniferous elements in Early Extreme Depositional Environments: Mega-End Members in Geologic Time:
Permian tropical floras, in Greb, S.F., and DiMichele, W.A., eds., Wetlands through Geological Society of America, Special Publication 370, p. 127–150.
Time: Geological Society of America, Special Paper 399, p. 223–248. GULBRANSON, E.L., MONTAÑEZ, I.P., SCHMITZ, M.D., LIMARINO, C.O., ISBELL, J.L.,
DIMICHELE, W.A., FALCON-LANG, H.J., NELSON, W.J., ELRICK, S.D., AND AMES, P.R., MARENSSI, S.A., AND CROWLEY, J.L., 2010, High-resolution U-Pb calibration of
2007, Ecological gradients within a Middle Pennsylvanian peat mire forest: Geology, Carboniferous glacigenic deposits, Rio Blanco and Paganzo basins, northwest
v. 35, p. 415–418. Argentina: Geological Society of America, Bulletin 122, p. 1480–1498.
DIMICHELE, W.D., MONTAÑEZ, I.P., POULSEN, C.J., AND TABOR, N.J., 2009, Feedbacks HAQ, B.U., AND SCHUTTER, S.R., 2008, A chronology of Paleozoic sea-level changes:
and regime shifts in the Late Palaeozoic ice-age Earth: Geobiology, v. 7, p. 200–226. Science, v. 322, p. 64–68.
DIMICHELE, W.A., CECIL, C.B., MONTAÑEZ, I.P., AND FALCON-LANG, H.J., 2010, Cyclic HECKEL, P.H., 1977, Origin of phosphatic black shale facies in Pennsylvanian
changes in Pennsylvanian paleoclimate and effects on floristic dynamics in tropical cyclothems of Mid-continent North America: American Association of Petroleum
Pangaea: International Journal of Coal Geology, v. 83, p. 329–344. Geologists, Bulletin, v. 61, p. 1045–1068.
DRIESE, S.G., AND OBER, E.G., 2005, Paleopedologic and paleohydrologic records of HECKEL, P.H., 1986, Sea-level curve for Pennsylvanian eustatic marine transgressive–
precipitation seasonality from Early Pennsylvanian ‘‘underclay’’ paleosols, U.S.A.: regressive depositional cycles along midcontinent outcrop belt, North America:
Journal of Sedimentary Research, v. 75, p. 997–1010. Geology, v. 14, p. 330–334.
666 N.A. ROSENAU ET AL. JSR
HECKEL, P.H., 1994, Evaluation of evidence for glacio-eustatic control over marine LAWRENCE, J.R., AND RASHKES-MEAUX, J., 1993, The stable isotopic composition of
Pennsylvanian cyclothems in North America and consideration of possible tectonic ancient kaolinites of North America, in Swart, P.K., Lohmann, K.C., McKenzie, J.,
effects, in Dennison, J.M., and Ettensohn, F.R., eds., Tectonic and Eustatic Controls and Savin, S., eds., Climate Change in Continental Isotopic Records: American
on Sedimentary Cycles: SEPM, Concepts in Sedimentology and Paleontology, v. 4, p. Geophysical Union, Geophysical Monograph 78, p. 249–262.
65–87. LAWRENCE, J.R., AND TAYLOR, H.P., 1971, Deuterium and 18O correlation: clay minerals
HECKEL, P.H., 2002, Overview of Pennsylvanian cyclothems in Midcontinent North and hydroxides in Quaternary soils compares to meteoric waters: Geochimica et
America and brief summary of those elsewhere in the world, in Hills, L.V., Cosmochimica Acta, v. 34, p. 25–42.
Hunderson, C.M., and Bamber, E.W., eds., The Carboniferous and Permian of the LAWRENCE, J.R., AND TAYLOR, H.P., 1972, Hydrogen and oxygen isotope systematics in
World: Canadian Society of Petroleum Geologists, Memoir 19, p. 79–98. weathering problems: Geochimica et Cosmochimica Acta, v. 36, p. 1377–1393.
HECKEL, P.H., 2008, Pennsylvanian cyclothems in Midcontinent North America as far- LECKIE, D.A., FOX, C., AND TAROCAI, C., 1989, Multiple paleosols of the Late Albian
field effects of waxing and waning of Gondwana ice sheets, in Fielding, C.R., Frank, Boulder Creek Formation, British Columbia, Canada: Sedimentology, v. 36, p. 307–
T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and Space: 323.
Geological Society of America, Special Paper 441, p. 275–289. LÉZINE, A.M., AND CHATEAUNEUF, J.J., 1991, Peat in the ‘‘Niayes’’ of Senegal:
HECKEL, P.H., AND WEIBEL, C. P., 1991, Current status of conodont-based depositional environment and Holocene evolution: Journal of African Earth Sciences,
biostratigraphic correlation of upper Pennsylvanian succession between Illinois and v. 12, p. 171–179.
midcontinent, in Weibel, C.P., ed., Sequence Stratigraphy in Mixed Clastic– LUDVIGSON, G.A., GONZALEZ, L.A., METZGER, R.A., WITZKE, B.J., BRENNER, R.L.,
Carbonate Strata, upper Pennsylvanian, East-Central Illinois: Champaign, SEPM, MURILLO, A.P., AND WHITE, T.S., 1998, Meteoric sphaerosiderite lines and their use
Great Lakes Section, 21st Annual Field Conference, Illinois State Geological Survey, for paleohydrology and paleoclimatology: Geology, v. 26, p. 1039–1042.
60–69. MCCARTHY, P.J., AND PLINT, A.G., 2003, Spatial variability of palaeosols across
HOPKINS, M.E., 1958, Geology and petrology of the Anvil Rock Sandstone of southern Cretaceous interfluves in the Dunvegan Formation, NE British Columbia, Canada:
Illinois: Illinois State Geological Survey, Circular 256, 49 p. palaeohydrological, palaeogeomorphological, and stratigraphic implications: Sedi-
HOPKINS, M.E., 1968, Harrisburg (No. 5) coal reserves of southeastern Illinois: Illinois mentology, v. 50, p. 1187–1220.
Geological Survey, Circular 431, 25 p. MCCARTHY, P.J., MARTINI, I.P., AND LECKIE, D.A., 1997, Pedosedimentary history and
HOPKINS, M.E., AND SIMON, J.A., 1975, Pennsylvanian System, in Handbook of Illinois floodplain dynamics of the Lower Cretaceous upper Blairmore Group, southwestern
Stratigraphy: Illinois Geological Survey, Bulletin 95, p. 163–201. Alberta, Canada: Canadian Journal of Earth Sciences, v. 34, p. 598–617.
HORTON, D.E., POULSEN, C.J., MONTAÑEZ, I.P., AND DIMICHELE, W.A., 2012, MCCARTHY, P.J., FACCINI, U.F., AND PLINT, A.G., 1999, Evolution of an ancient coastal
Eccentricity-paced late Paleozoic climate change: Palaeogeography, Palaeoclimatol- plain: palaeosols, interfluves and alluvial architecture in a sequence stratigraphic
ogy, Palaeoecology, v. 331–332, p. 150–161. framework, Cenomanian Dunvegan Formation, NE British Columbia, Canada:
HSIEH, J.C.C., 1997, An oxygen isotopic study of soil water and pedogenic clays in Sedimentology, v. 46, p. 861–891.
Hawaii [unpublished Ph.D. Thesis]: California Institute of Technology, 181 p. MCCREA, J.M., 1950, On the isotopic chemistry of carbonates and a paleotemperature
IRWIN, H., CURTIS, C., AND COLEMAN, M., 1977, Isotopic evidence for source of scale: Journal of Chemical Physics, v. 18, p. 849–857.
diagenetic carbonates formed during burial of organic-rich sediments: Nature, v. 269, MILLER, D.C., AND BIRKELAND, P.W., 1992, Soil catena variation along an alpine
p. 209–213. climatic transect, northern Peruvian Andes: Geoderma, v. 55, p. 211–223.
ISBELL, J.L., MILLER, M.F., WOLFE, K.L., AND LENAKER, P.A., 2003, Timing of the late MILLER, K.B., AND WEST, R.R., 1993, A Reevaluaation of Wolfcampian Cyclothems in
Paleozoic glaciation in Gondwana: Was glaciation responsible for the development of Northeastern Kansas: Significance of Subaerial Exposure and Flooding Surfaces:
northern hemisphere cyclothems? in Chan, M.A., and Archer, A.W., eds., Extreme Kansas Geological Survey, Bulletin 235, p. 1–26.
Depositional Environments: Mega End Members in Geologic Time: Geological MILLER, K.B., MCCAHON, T.J., AND WEST, R.R., 1996, Lower Permian (Wolfcampian)
Society of America, Special Publication 370, p. 5–24. paleosol-bearing cycles of the U.S. Midcontinent: evidence of climatic cyclicity:
JENNY, H., 1980, The Soil Resource: Origin and Behavior: New York, Springer, Journal of Sedimentary Research, v. 66, p. 71–84.
Ecological Studies, 37, 377 p. MONTAÑEZ, I.P., TABOR, N.J., NIEMEIR, D., DIMICHELE, W.A., FRANK, T.D., FIELDING,
C.R., ISBELL, J.L., BIRGENHEIER, L.P., AND RYGEL, M.C., 2007, CO2-forced climate
JOECKEL, R.M., 1991, Paleosol stratigraphy of the Eskridge Formation: early Permian
and vegetation instability during late Paleozoic deglaciation: Science, v. 315, p. 87–91.
pedogenesis and climate in southeastern Nebraska: Journal of Sedimentary Petrology,
v. 61, p. 234–255. MOORE, D.M., AND REYNOLDS, R.C., 1997, X-Ray Diffraction and the Identification and
Analysis of Clay Minerals: New York, Oxford University Press, 378 p.
JOECKEL, R.M., 1995, Paleosols below the Ames Marine Unit (Upper Pennsylvanian,
MOZLEY, P.S., 1989, Relation between depositional environment and the elemental
Conemaugh Group) in the Appalachian Basin, USA: variability on an ancient
composition of early diagenetic siderite: Geology, v. 17, p. 704–706.
depositional landscape: Journal of Sedimentary Research, v. 65, p. 393–407.
MOZLEY P.S., AND BURNS, S.J., 1993, Oxygen and carbon isotopic composition of marine
KELLER, W.D., 1968, Flint clay and flint clay facies: Clays and Clay Minerals, v. 16, p.
carbonate concretions: an overview: Journal of Sedimentary Petrology, v. 63, p. 73–
113–128.
83.
KELLER, W.D., 1975, Refractory clay in the lower part of the Pennsylvanian System:
MOZLEY, P.S., AND CAROTHERS, W.W., 1992, Elemental and isotopic composition of
U.S. Geological Survey, Professional Paper 853, p. 65–71. siderite in the Daruk Formation, Alaska: effect of microbial activity and water–
KELLER, W.D., 1981, The sedimentology of flint clay: Journal of Sedimentary Petrology, sediment interaction on early pore-water chemistry: Journal of Sedimentary
v. 51, p. 233–244. Petrology, v. 62, p. 681–692.
KLEIN, G.D., AND CLOETINGH, S., 1989, Tectonic subsidence during deposition of NELSON, W.J., 1983, Geologic Disturbances in Illinois Coal Seams: Illinois State
Pennsylvanian cyclothems [abstract]: Washington, D.C., 28th International Geolog- Geological Survey, Circular 530, 47 p.
ical Congress, Abstracts, v. 2, p. 198–199. NELSON, W.J., ELRICK, S.D., AND AMES, P., 2010, ‘‘Merging’’ Pennsylvanian limestones
KLEIN, G.D., AND KUPPERMAN, J.B., 1992, Pennsylvanian cyclothems: methods of in Illinois basin: formation at highstand [abstract]: Geological Society of America,
distinguishing tectonically induced changes in sea level from climatically induced Abstracts with Programs, v. 42, no. 2, p. 51.
changes: Geological Society of America, Bulletin, 104, p. 166–175. NELSON, W.J., ELRICK, S.D., AND ASKARI KHORASGANI, Z., 2011, Revised Upper
KLEIN, G.D., AND WILLARD, D.A, 1989, Origin of the Pennsylvanian coal- bearing Pennsylvanian stratigraphy of Illinois Basin with regional correlations [abstract]:
cyclothems of North America: Geology, v. 17, p. 152–155. Geological Society of America, Abstracts with Programs, v. 43, no. 5, p. 603.
KOSANKE, R.M., AND CECIL, C.B., 1996, Late Pennsylvanian climate changes and OLSZEWSKI, T.D., AND PATZKOWSKY, M.E., 2003, From cyclothems to sequences: the
palynomorph extinctions: Review of Palaeobotany and Palynology, v. 90, p. 113–140. record of eustasy and climate on an icehouse epeiric platform (Pennsylvanian–
KRAUS, M.J., 1987, Integration of channel and floodplain suites: II. Lateral relations of Permian, North American Midcontinent): Journal of Sedimentary Research, v. 73, p.
alluvial paleosols: Journal of Sedimentary Petrology, v. 57, p. 602–612. 15–30.
KRAUS, M.J., AND ASLAN, A., 1993, Eocene hydromorphic paleosols: significance for PARRISH, J.T., 1993, Climate of the supercontinent Pangaea: Journal of Geology, v. 101,
interpreting ancient floodplain processes: Journal of Sedimentary Petrology, v. 63, p. p. 215–233.
453–463. PATTERSON, S.H., AND HOSTERMAN, J.W., 1960, Geology of the clay deposits in the Olive
KRAUS, M.J., AND ASLAN, A., 1999, Paleosol sequences in floodplain environments: a Hill district, Kentucky, in Swineford, A., ed., Clays and Clay Minerals: Proceedings of
hierarchical approach, in Thiry, M.Z., ed., Palaeoweathering, Palaeosurfaces and the Seventh National Conference on Clays and Clay Minerals: Washington, D.C.,
Related Continental Deposits: International Association of Sedimentologists, Special 1958, New York, Pergamon Press, p. 178–194.
Publication 27, p. 303–321. PEPPERS, R.A., 1964, Spores in strata of Late Pennsylvanian cyclothems in the Illinois
KRAUS, M.J., AND BOWN, T.M., 1993, Short-term sediment accumulation rates Basin: Illinois State Geological Survey, Bulletin 90, p. 1–89.
determined from Eocene alluvial paleosols: Geology, v. 21, p. 743–746. PEPPERS, R.A., 1996, Palynological Correlation of Major Pennsylvanian (Middle and
KVALE, E.P., FRASER, G.S., ARCHER, A.W., ZAWISTOSKI, A., KEMP, N., AND MCGOUGH, Upper Carboniferous) Chronostratigraphic Boundaries in the Illinois and Other Coal
P., 1994, Evidence of seasonal precipitation in Pennsylvanian sediments of the Illinois Basins: Geological Society of America, Memoir 188, 111 p.
Basin: Geology, v. 22, p. 331–334. PEPPERS, R.A., 1997, Palynology of the Lost Branch Formation of Kansas: new insights
LAHTEENOJA, O., AND ROUCOUX, K.H., 2010, Inception, history, and development of on the major floral transition at the Middle–Upper Pennsylvanian boundary: Review
peatlands in the Amazon Basin: Pages News, v. 18, p. 27–31. of Palaeobotany and Palynology, v. 98, p. 223–246.
LANGENHEIM, R., AND NELSON, W., 1992, The cyclothemic concept in the Illinois Basin: a PERLMUTTER, M.A., AND MATTHEWS, M.D., 1989, Global cyclostratigraphy —a model,
review, in DOTT, R.H., JR., Eustasy: The Historical Ups and Downs of a Major in Cross, T.A., ed., Quantitative Dynamic Stratigraphy: Englewood Cliffs, New
Geological Concept: Geological Society of America, Memoir 180, p. 55–71. Jersey, Prentice Hall, p. 223–260.
JSR PENNSYLVANIAN PALEOSOLS OF THE ILLINOIS BASIN: PART II 667
PEYSER, C.E., AND POULSEN, C.J., 2008, Controls on Permo-Carboniferous precipitation SOREGHAN, G.S., 2007, Reconciling indicators of cold and warmth in Late Palaeozoic
over tropical Pangaea: A GCM sensitivity study: Palaeogeography, Palaeoclimatol- tropical Pangaea: high-magnitude glacial–interglacial climate change? [abstract]:
ogy, Palaeoecology, v. 268, p. 181–192. Geological Society of America, Annual Convention, Abstracts with Programs, v. 38,
PFEFFERKORN, H.W., AND THOMSON, M.C., 1982, Changes in dominance patterns in p. 36.
Upper Carboniferous plant fossil assemblages: Geology, v. 10, p. 641–644. SOREGHAN, G.S., AND GILES, K.A., 1999, Amplitudes of Late Pennsylvanian
PFEFFERKORN, H.W., GASTALDO, R.A., DIMICHELE, W.A., AND PHILLIPS, T.L., 2008, glacioeustasy: Geology, v. 27, p. 255–258.
Pennsylvanian tropical floras from the United States as a record of changing climate, SOREGHAN, G.S., SOREGHAN, M.J., AND EBLE, C.F., 2002a, How high were the Ancestral
in Fielding, C.R., Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Rocky Mountains? [abstract]: Geological Society of America, Annual Convention,
Age in Time and Space: Geological Society of America, Special Paper 441, p. 305–316. Abstracts with Programs, v. 34, p. 62.
PHILLIPS, T.L., 1979, Reproduction of heterosporous arborescent lycopods in the SOREGHAN, M.J., SOREGHAN, G.S., AND HAMILTON, M.A., 2002b, Palaeo-winds inferred
Mississippian–Pennsylvanian of Euramerica: Review of Palaeobotany and Palynol- from detrital-zircon geochronology of upper Palaeozoic loessite, western equatorial
ogy, v. 27, p. 239–289. Pangaea: Geology, v. 30, p. 695–698.
PHILLIPS, T.L., AND PEPPERS, R.A., 1984, Changing patterns of Pennsylvanian coal- STAUB, J.R., AND ESTERLE, J.S., 1994, Peat-accumulating depositional systems of
swamp vegetation and implications of climatic control on coal occurrence: Sarawak, East Malaysia: Sedimentary Geology, v. 89, p. 91–106.
International Journal of Coal Geology, v. 3, p. 205–255. STAUB, J.R., AND GASTALDO, R.A., 2003, Late Quaternary sedimentation and peat
PHILLIPS, T.L., PEPPERS, R.A., AVCIN, M.J., AND LAUGHNAN, P.F., 1974, Fossil plants and development in the Rajang River Delta, Sarawak, East Malaysia in Sidi, F.H. et al.,
coal: patterns of change in Pennsylvanian coal swamps of the Illinois Basin: Science, v. Tropical deltas of southeast Asia: sedimentology, stratigraphy, and petroleum
184, p. 1367–1369. geology: SEPM, Special Publication 76, p. 71–87.
PHILLIPS, T.L., PEPPERS, R.A., AND DIMICHELE, W.A., 1985, Stratigraphic and STERN, L.A., CHAMBERLAIN, C.P., REYNOLDS, R.C., AND JOHNSON, G.D., 1997, Oxygen
interregional changes in Pennsylvanian coal swamp vegetation: environmental isotope evidence of climate change from pedogenic clay minerals in the Himalayan
inferences: International Journal of Coal Geology, v. 5, p. 43–109. molasse: Geochimica et Cosmochimica Acta, v. 61, p. 731–744.
PLINT, A.G., MCCARTHY, P.J., AND FACCINI, U.F., 2001, Nonmarine sequence SUAREZ, M.B., GONZALEZ, L.A., AND LUDVIGSON, G.A., 2010, Estimating the oxygen
stratigraphy: updip expression of sequence boundaries and systems tracts in a high- isotopic composition of equatorial precipitation during the mid-Cretaceous: Journal
resolution framework, Cenomanian Dunvegan Formation, Alberta foreland basin, of Sedimentary Research, v. 80, p. 480–491.
Canada: American Association of Petroleum Geologists, Bulletin, v. 85, p. 1967–2001. TABOR, N.J., AND MONTAÑEZ, I.P., 2002, Shifts in late Paleozoic atmospheric circulation
POSTMA, D., 1981, Formation of siderite and vivianite and the pore-water composition over western equatorial Pangaea: insights from pedogenic mineral d18O compositions:
of a recent bog sediment in Denmark: Chemical Geology, v. 31, p. 225–244. Geology, v. 30, p. 1127–1130.
POSTMA, D., 1982, Pyrite and siderite formation in brackish and freshwater swamp TABOR, N.J., AND MONTAÑEZ, I.P., 2004, Morphology and distribution of fossil soils in
sediments: American Journal of Science, v. 282, p. 1151–1183. the Permo-Pennsylvanian Wichita and Bowie groups, north-central Texas, USA:
POTTER, P.E., 1962, Sand Body Shape and Map Pattern of Pennsylvanian Sandstones in implications for western equatorial Pangaean palaeoclimate during icehouse–
Illinois: Illinois State Geological Survey, Circular 339, 36 p. greenhouse transition: Sedimentology, v. 51, p. 851–884.
POTTER, P.E., 1963, Late Paleozoic sandstones of the Illinois Basin: Illinois State TABOR, N.J., AND MONTAÑEZ, I.P., 2005, Oxygen and hydrogen isotope compositions of
Geological Survey, Report of Investigations 217, 92 p. and 9 plates. Permian pedogenic phyllosilicates: development of modern surface domain arrays and
POULSEN, C.J., POLLARD, D., MONTAÑEZ, I.P., AND ROWLEY, D., 2007, Late Palaeozoic implications for paleotemperature reconstruction: Palaeogeography, Palaeoclimatol-
tropical climate response to Gondwana deglaciation: Geology, v. 35, p. 771–774. ogy, Palaeoecology, v. 223, p. 127–146.
QUADE, J., CERLING, T.E., AND BOWMAN, J.R., 1989, Development of Asian monsoon TABOR, N.J., AND POULSEN, C.J., 2008, Palaeoclimate across the Late Pennsylvanian–
revealed by marked ecological shift during the latest Miocene in northern Pakistan: Early Permian tropical palaeolatitudes: a review of climate indicators, their
Nature, v. 342, p. 162–166. distribution, and relation to palaeophysiographic climate factors: Palaeogeography,
Palaeoclimatology, Palaeoecology, v. 268, p. 293–310.
QUINLAN, G.M., AND BEAUMONT, C., 1984, Appalachian thrusting, lithospheric flexure,
and the Paleozoic stratigraphy of the eastern interior of North America: Canadian TABOR, N.J., AND YAPP, C.J., 2005, Coexisting goethite and gibbsite from a high-
Journal of Earth Sciences, v. 21, p. 973–996. paleolatitude (55uN) Late Paleocene laterite: concentration and 13C/12C ratios of
occluded CO2 and associated organic matter: Geochimica et Cosmochimica Acta, v.
RETALLACK, G.J., 1983, A paleopedological approach to the interpretation of terrestrial
69, p. 5495–5510.
sedimentary rocks: the mid Tertiary fossil soils of Badlands National Park, South
TABOR, N.J., MONTAÑEZ, I.P., AND SOUTHARD, R.J., 2002, Paleoenvironmental
Dakota: Geological Society of America, Bulletin, v. 94, p. 823–840.
reconstruction from chemical and isotopic compositions of Permo-Pennsylvanian
ROSENAU, N.A., TABOR, N.J., ELRICK, S.E, AND NELSON, W.J., 2013, Polygenetic history
pedogenic minerals: Geochimica et Cosmochimica Acta, v. 66, p. 3093–3107.
of paleosols in middle–upper Pennsylvanian cyclothems of the Illinois basin, USA:
TABOR, N.J., MONTAÑEZ, I.P, ZIERENBERG, R., AND CURRIE, B.S., 2004, Mineralogical
Part I. Characterization of paleosol types and interpretation of pedogenic processes:
and geochemical evolution of a basalt-hosted fossil soil (Late Triassic, Ischigualasto
Journal of Sedimentary Research, v. 83, p. xxx–xxx.
Formation, northwest Argentina): potential for paleoenvironmental reconstruction:
ROSSCOE, S.R., 2008, Idiognathodus and Streptognathodus species from the Lost Branch Geological Society of America, Bulletin, v. 116, p. 1280–1293.
to Dewey sequences (Middle–Upper Pennsylvanian) of the Midcontinent Basin, TABOR, N.J., MONTAÑEZ, I.P., KELSO, K.A., CURRIE, B.S., AND SHIPPMAN, T.A., 2006, A
North America [Unpublished Ph.D. thesis]: Lubbock, Texas, Texas Tech University, late Triassic soil catena: landscape controls on paleosol morphology across the
191 p. Carnian-Age Ischigualasto–Villa Union Basin, Northwestern Argentina, in Alonso-
ROWLEY, D.B., RAYMOND, A., PARRISH, J.T., LOTTES, A.L., SCOTESE, R., AND ZIEGLER, Zarza, A.M., and Tanner, L.H., eds., Paleoenvironmental Record and Applications of
A.M., 1985, Carboniferous paleogeographic, phytogeographic and paleoclimatic Calcretes and Palustrine Carbonates: Geological Society of America, Special Paper
reconstructions: International Journal of Coal Geology, v. 5, p. 7–42. 416, p. 17–41.
ROZANSKI, K., ARAGUAS-ARAGUAS, L., AND GONFIANTINI, R., 1993, Isotopic patterns in TABOR, N.J., MONTAÑEZ, I.P., SCOTESE, C.R., POULSEN, C.J., AND MACk, G.H., 2008,
modern global precipitation, in McKenzie, J.A., and Savin, S., eds., Climate Change Paleosol archives of environmental and climatic history in paleotropical western
in Continental Isotopic Records: American Geophysical Union, Geophysical Pangaea during the latest Pennsylvanian through Early Permian, in Fielding, C.R.,
Monograph 78, p. 1–36. Frank, T.D., and Isbell, J.L., eds., Resolving the Late Paleozoic Ice Age in Time and
RYGEL, M.C., FIELDING, C.R., FRANK, T.D., AND BIRGENHEIER, L.P., 2008, The Space: Geological Society of America, Special Paper 441, p. 291–303.
magnitude of Late Paleozoic glacioeustatic fluctuations: a synthesis: Journal of TANDON, S.K., AND GIBLING, M.R., 1994, Calcrete and coal in Late Carboniferous
Sedimentary Research, v. 78, p. 500–511. cyclothems of Nova Scotia, Canada: climate and sea-level changes linked: Geology, v.
SAHNEY, S., BENTON, M.J., AND FALCON-LANG, H.J., 2010, Rainforest collapse triggered 22, p. 755–758.
Carboniferous tetrapod diversification in Euramerica: Geology, v. 38, p. 1079–1082. TANKARD, A.J., 1986, Depositional response to foreland deformation in the
SAVIN, S.M., AND EPSTEIN, S., 1970, The oxygen and hydrogen isotope geochemistry of Carboniferous of eastern Kentucky: American Association of Petroleum Geologists,
clay minerals: Geochimica et Cosmochimica Acta, v. 34, p. 25–42. Bulletin, v. 70, p. 853–868.
SHELDON, N.D., AND TABOR, N.J., 2009, Quantitative paleoenvironmental and UFNAR, D.F., GONZALEZ, L.A., LUDVIGSON, G.A., BRENNER, R.L., AND WITZKE, B.J.,
paleoclimatic reconstruction using paleosols: Earth-Science Reviews, v. 95, p. 1–52. 2001, Stratigraphic implications of meteoric sphaerosiderite d18O values in paleosols
SHEPPARD, S.F.M., AND GILG, H.A., 1996, Stable isotope geochemistry of clay minerals: in the Cretaceous (Albian) Boulder Creek Formation, NE British Columbia Foothills,
Clay Minerals, v. 31, p. 1–24. Canada: Journal of Sedimentary Research, v. 71, p. 1017–1028.
SIEVER, R., 1957, Pennsylvanian sandstones of the Eastern Interior coal basin: Journal UFNAR, D.F., GONZÁLEZ, L.A., LUDVIGSON, G.A., BRENNER, R.L., AND WITZKE, B.J.,
of Sedimentary Petrology, v. 27, p. 227–250. 2002, The mid-Cretaceous water bearer: isotope mass balance quantification of the
SMITH, R.M.H., 1990, Alluvial paleosols and pedofacies sequences in the Permian Lower Albian hydrologic cycle: Palaeogeography, Palaeoclimatology, Palaeoecology, v. 188,
Beaufort of the southwestern Karoo Basin, South Africa: Journal of Sedimentary p. 51–71.
Petrology, v. 60, p. 258–276. UFNAR, D.F., LUDVIGSON, G.A., GONZÁLEZ, L.A., BRENNER, R.L., AND WITZKE, B.J.,
SMITH, W.H., AND O’BRIEN, N.R., 1965, Middle and late Pennsylvanian flint clays: 2004, High latitude meteoric 18O compositions: paleosol siderite in the middle
Journal of Sedimentary Petrology, v. 35, p. 610–618. Cretaceous Nanushuk Formation, North Slope, Alaska: Geological Society of
SOREGHAN, G.S., 1994, The impact of glacioclimatic change on Pennsylvanian America, Bulletin, v. 16, p. 463–473.
cyclostratigraphy, in Embry, A.F., Beauchamp, B., and Glass, D.J., eds., VITALI, F., LONGSTAFFE, F.J., BIRD, M.I., AND CALDWELL, W.G.E., 2000, Oxygen-isotope
Pangaea—Global Environments and Resources: Canadian Society of Petroleum fraction between aluminum-hydroxide phases and water at , 60uC: results of decade-
Geologist, Memoir 17, p. 523–554. long synthesis experiments: Clay and Clay Minerals, v. 48, p. 230–237.
668 N.A. ROSENAU ET AL. JSR
VITALI, F., LONGSTAFFE, F.J., BIRD, M.I., GAGE, K.L., AND CALDWELL, W.G.E., 2001, WILSON, M.J., 1999, The origin and formation of clay minerals in soils: past, present and
Hydrogen-isotope fractionation in aluminum hydroxides: synthesis products versus future perspectives: Clay Minerals, v. 34, p. 7–25.
natural samples from bauxites: Geochimica et Cosmochimica Acta, v. 65, p. 1391– WRIGHT, V. P., AND MARRIOTT, S.B., 1993, The sequence stratigraphy of fluvial
1398. depositional systems: the role of floodplain sediment storage: Sedimentary Geology, v.
VITALI, F., LONGSTAFFE, F.J., MCCARTHY, P.J., PLINT, A.G., AND CALDWELL, W.G.E., 86, p. 203–210.
2002, Stable isotopic investigation of clay minerals and pedogenesis in an interfluve WRIGHT, V.P., AND ROBINSON, D., 1988, Early Carboniferous floodplain deposits from
paleosol from the Cenomanian Dunvegan Formation, N.E. British Columbia, South Wales: a case study of the controls on palaeosol development: Geological
Canada: Chemical Geology, v. 192, p. 269–287. Society of London, Journal, v. 145, p. 847–857.
WAAGE, K., 1950, Refractory Clays of the Maryland Coal Measures: Maryland YAALON, D.H., AND KALMAR, D., 1978, Dynamics of cracking and swelling clay soils:
Department of Geology, Mines and Water Resources Bulletin 9, 182 p. displacement of skeletal grains, optimum depth of slickensides, and rate of intra–
WANLESS, H.R., AND SHEPARD, F.P., 1936, Sea level and climatic changes related to late pedonic turbation: Earth Surface Processes and Landforms, v. 3, p. 31–42.
Paleozoic cycles: Geological Society of America, Bulletin, v. 47, p. 1177–1206. YAPP, C.J., 1993, The stable isotope geochemistry of low temperature Fe(III) and Al
WANLESS, H.R., AND WELLER, J.M., 1932, Correlation and extent of Pennsylvanian ‘‘Oxides’’ with implications for continental paleoclimates, in Swart, P.R., Lohmann,
cyclothems: Geological Society of America, Bulletin, v. 43, p. 1003–1016. KC, McKenzie, J., and Savin, S., eds., Climate Change in Continental Isotopic
WELLER, J.M., 1930, Cyclical sedimentation of the Pennsylvanian Period and its Records: American Geophysical Union, Geophysical Monograph 78, p. 285–294.
significance: Journal of Geology, v. 38, p. 97–135. YAPP, C.J., 2001a, Mixing of CO2 in surficial environments as recorded by the
WELLER, J.M., 1931, The conception of cyclic sedimentation during the Pennsylvanian concentration and d13C values of the Fe(CO3)OH component in goethite: Geochimica
Period: Illinois Geological Survey, Bulletin 60, p. 163–177. et Cosmochimica Acta, v. 65, p. 4115–4130.
WELLER, J.M., 1956, Argument for diastrophic control of late Paleozoic cyclothems: YAPP, C.J., 2001b, Rusty relics of earth history: Iron (III) oxides, isotopes, and surficial
American Association of Petroleum Geologists, Bulletin, v. 40, p. 17–50. environments: Annual Review of Earth and Planetary Sciences, v. 29, p. 165–199.
WEST, R.R., ARCHER, A.W., AND MILLER, K.B., 1997, The role of climate in stratigraphic YAPP, C.J., 2002, Erratum to Crayton J. Yapp (2001), ‘‘Mixing of CO2 in surficial
patterns exhibited by late Paleozoic rocks exposed in Kansas: Palaeogeography, environments as recorded by the concentration and d13C values of the Fe(CO3)OH
Palaeoclimatology, Palaeoecology, v. 128, p. 1–16. component in goethite,’’ GCA, Vol. 65, 4115–4130: Geochimica et Cosmochimica
Acta, v. 66, p. 497.
WHITE, D., 1913, Climates of coal forming periods, in White, C.D., and Theissen, R.,
YEMANE, K., KAHR, G., AND KELTS, K., 1996, Imprints of post-glacial climates and
eds., The Origin of Coal: U.S. Bureau of Mines, Bulletin, v. 38, p. 68–79.
paleogeography in the detrital clay mineral assemblages of an Upper Permian fluvio-
WHITE, D., 1925, Environmental conditions of deposition of coal: American Institute of lacustrine Gondwana deposit from north-central Malawi: Palaeogeography, Palaeo-
Mining and Metallurgical Engineers, Transactions, v. 71, p. 3–34. climatology, Palaeoecology, v. 125, p. 27–49.
WHITE, T.S, GONZÁLEZ, L.A, LUDVIGSON, G.A., AND POULSEN, C.J., 2001, Middle ZANGERL, R., AND RICHARDSON, E.S., 1963, The paleoecological history of two
Cretaceous greenhouse hydrologic cycle of North America: Geology, v. 29, p. 363– Pennsylvanian black shales: Fieldiana, Geology Memoir 4, 352 p.
366. ZIEGLER, A.M., RAYMOND, A.A., GEIRLOWSKI, T.C., HORRELL, M.A., ROWLEY, D.B., AND
WHITICAR, M.J., FABER, E., AND SCHOELL, M., 1986, Biogenic methane formation in LOTTES, A.L., 1987, Coal, climate and terrestrial productivity: the present and early
marine and freshwater environments: CO2 reductions vs. acetate fermentation- Cretaceous compared, in Scot, A.C., ed., Coal and Coal-bearing Strata: Recent
isotope evidence: Geochimica et Cosmochimica Acta, v. 50, p. 693–709. Advances: Geological Society of London, Special Paper 32, p. 25–49.
WILDING, L.P., AND TESSIER, D., 1988, Genesis of Vertisols: shrink–swell phenomena, in ZIEGLER, A.M., HULVER, M.L., AND ROWLEY, D.B., 1997, Permian world topography
Wilding, L.P., and Puentes, R., eds., Vertisols: Their Distribution, Properties, and climate, in Martinit, I.P., ed., Late Glacial and Post-Glacial Environmental
Classification and Management: College Station, Texas, Texas A&M University, Changes: Quaternary, Carboniferous–Permian and Proterozoic: Oxford, U.K.,
Publishing Center, p. 55–81. Oxford University Press, p. 111–146.
WILLIAMS, E.G., 1960, Relationship between the stratigraphy and petrography of ZIEGLER, A.M., REES, P.M., AND NAUGOLNYKH, S.V., 2002, The Early Permian floras of
Pottsville sandstones and the occurrence of high-alumina Mercer clay: Economic Prince Edward Island, Canada: differentiating global from local effects of climate
Geology, v. 55, p. 1291–1302. change: Canadian Journal of Earth Sciences, v. 39, p. 223–238.
WILLIAMS, E.G., BERBENBACK, R.F., FALLA, W.S., AND UDAGAWA, S., 1968, Origin of
some Pennsylvanian underelays in western Pennsylvania: Journal of Sedimentary
Petrology, v. 39, p. 1179–1193. Received 30 October 2012; accepted 16 April 2013.