Você está na página 1de 29

Cretaceous Research 56 (2015) 316e344

Contents lists available at ScienceDirect

Cretaceous Research
journal homepage: www.elsevier.com/locate/CretRes

An astronomically calibrated stratigraphy of the Cenomanian,


Turonian and earliest Coniacian from the Cretaceous Western Interior
Seaway, USA: Implications for global chronostratigraphy
James S. Eldrett a, *, Chao Ma a, b, Steven C. Bergman a, Brendan Lutz a, 1, F. John Gregory c,
Paul Dodsworth d, Mark Phipps c, Petros Hardas c, Daniel Minisini a, Aysen Ozkan a, 1,
Jahander Ramezani e, Samuel A. Bowring e, Sandra L. Kamo f, Kurt Ferguson g,
Calum Macaulay a, Amy E. Kelly a
a
Shell International Exploration and Production Inc, 3333 Highway 6 South, Houston, TX 77082, USA
b
Department of Geoscience, University of Wisconsin-Madison, 1215 West Dayton Street, Madison, WI 53706, USA
c
Petrostrat Ltd., Tan-y-Graig, Parc Caer Seion, Conwy, North Wales, LL32 8FA, UK
d
StrataSolve Ltd, 42 Gaskell Street, Stockton Heath, Warrington, WA4 2UN, UK
e
Department of Earth, Atmospheric and Planetary Sciences, Massachusetts Institute of Technology, 77 Massachusetts Avenue, Cambridge, MA 02139, USA
f
Jack Satterly Geochronology Laboratory, Department of Geology, University of Toronto, 22 Russell St, Toronto, ON M5S 3B1, Canada
g
Hufngton School of Earth Sciences, Southern Methodist University, Dallas, TX 75275, USA

a r t i c l e i n f o a b s t r a c t

Article history: This study describes integrated, astronomically tuned age models for a relatively expanded section of the
Received 13 August 2014 Eagle Ford Group (Texas, USA) from the Shell Iona-1 research core, which encompass >8 Myr ranging
Accepted in revised form 20 April 2015 from the earliest Cenomanian to the earliest Coniacian. Biostratigraphy combined with UePb dates from
Available online 18 June 2015
individual zircons representing ten bentonites provide geochronologic constraints for the astronomical
analyses. The astronomically tuned age models were used to calibrate a full suite of regional and globally
Keywords:
recognized age diagnostic biostratigraphic (integrated micropaleontological, nannopaleontological and
Biostratigraphy
palynological analyses), geochemical, as well as isotopic events. Newly developed integrated astro-
Astrochronology
Geochronology
nomical age models provide two estimates for the Cenomanian-Turonian boundary age of 94.10 0.13
Isotope stratigraphy Ma and 94.07 0.16 Ma similar to that previously proposed from the Cretaceous Western Interior
Cenomanian-Turonian Seaway. The duration of Oceanic Anoxic Event-2 (OAE-2) was calcuated as 0.71 0.17 Myr, consistent
Eagle Ford Group with other records from the KWIS, and several globally correlatetable, precursor events prior to the main
Buda limestone CIE have been identied. Elsewhere in North America, the middle-upper Turonian strata are generally
Boquillas Formation missing due to a hiatus, whereas the new data indicate it is relatively intact in the Iona-1 core, and thus
Austin Chalk provides a critical record to ll this gap in our understanding. As such, the Iona-1 core is one of the most
complete and best-preserved records of the Cenomanian, Turonian and early Coniacian stages, and
therefore provides insights into Late Cretaceous paleoclimate and paleoenvironment during this critical
period in Earth history.
2015 Elsevier Ltd. All rights reserved.

1. Introduction in the carbon cycle termed Oceanic Anoxic Events (OAE's), the most
prominent occurring at the Cenomanian-Turonian (C-T) transition
The Late Cretaceous was characterized by sustained global (OAE-2; sensu Schlanger and Jenkyns, 1976). This interval in
warming, emplacement of Large Igneous Provinces (LIPs), global particular was characterized by the widespread deposition of
extinctions, epicontinental seaways, and major global perturbations organic-rich, ne-grained sediment marked by a globally recog-
nized positive carbon isotope excursion (CIE) reecting the wide-
* Corresponding author. spread removal of 12C-enriched organic matter in marine sediments
E-mail address: james.eldrett@shell.com (J.S. Eldrett). under global anoxic conditions (see Jenkyns, 2010 and references
1
Shell Exploration and Production Inc., 150 North Dairy Ashford Street, Houston, therein). However, the exact timing of this inferred global OAE-2
TX 77079, USA.

http://dx.doi.org/10.1016/j.cretres.2015.04.010
0195-6671/ 2015 Elsevier Ltd. All rights reserved.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 317

event is still debated with recent studies showing diachroneity


between the deposition of the organic-rich sediment and the CIE
(e.g., Sageman et al., 1997; Tsikos et al., 2004; Meyers et al., 2005;
Meyers, 2007; Trabucho Alexandre et al., 2010; Eldrett et al.,
2014), whereas interpretations vary on the precise timing and
origin of the trigger (Mitchell et al., 2008; Turgeon and Creaser,
2008; Du Vivier et al., 2014; Eldrett et al., 2014; Laurin et al., 2015).
To understand the underlying causal mechanisms of OAE-2 and
to accurately calculate the rate of biogeochemical processes, robust
and globally consistent high-resolution chronostratigraphic
frameworks are required. Astrochronology is a proven tool for
construction of high-resolution geological time scales using a va-
riety of stratigraphic characteristics and well-established Milan-
kovitch orbital cycles of eccentricity, obliquity, and precession (e.g.,
Hinnov and Hilgen, 2012), and consequently cyclostratigraphic
analyses of various OAE 2 sections have been conducted using a
Fig. 1. Late Cenomanian palaeogeographic map of the Cretaceous Western Interior
variety of proxies (e.g., Kuhnt et al., 1997, 2004; Caron et al., 1999; Seaway (KWIS) and the margin of the Gulf of Mexico Carbonate Shelf, showing the
Meyers et al., 2001, 2012a,b; Prokoph et al., 2001; Kuypers et al., location of Iona-1, Mustang-1 and USGS Portland-1 cores. SB: Sabinas Basin; MB:
2004; Sageman et al., 2006; Mitchell et al., 2008; Voigt et al., Maverick Basin; ETB: East Texas Basin. Adapted from Ron Blakey and the Colorado
2008; Lanci et al., 2010; Ma et al., 2014). However, these astro- Plateau Geosystems Inc.
chronologies are generally stratigraphically limited due to un-
conformities or partial sections with no single continuous,
2.2. Mustang Ranch-1
integrated coverage of the entire Cenomanian and Turonian stages
(Hinnov and Hilgen, 2012). Therefore, most of the analyses have
To supplement data from the Iona-1 core, comparable sedi-
focused on the OAE-2 interval (<1 Myr duration) and are incom-
mentary rocks were also analyzed for organic and inorganic carbon
plete because of missing stratigraphic records or poor signal to
isotopes and biostratigraphy (micropaleontology, nannopaleontol-
noise ratio. Here, integrated astronomically tuned age models are
ogy, palynology) from the Mustang Ranch-1 core (28 46.550 N, 99
presented for the Eagle Ford Group (Gr., Texas, USA) which en-
57.510 W) located in the Maverick Basin (Texas, USA; Fig. 1).
compasses >8 Myr of the earliest Cenomanian to the earliest
Coniacian stages and provides an exceptionally well-preserved
record. The astronomically tuned age models are also used to 2.3. USGS Portland-1
calibrate a full suite of regional and globally recognized age diag-
nostic biostratigraphic (integrated micropaleontological, nanno- Sedimentary rocks were also analyzed for d13Corg and biostra-
paleontological and palynological analyses), geochemical, and tigraphy (micropaleontology, palynology) from the USGS Portland-
isotopic events. 1 core (38 22.360 N, 105 01.180 W) located in the central KWIS, near
the GSSP section (Kennedy et al., 2000, 2005) at Pueblo (Colorado,
USA). The new data cover the Cenomanian and Turonian equiva-
2. Material and methods lents to the interval spanning the Dakota Sandstone, Graneros
Shale, Lincoln Limestone, Harland Shale and lowermost part of the
2.1. Shell research core Iona-1 Bridge Creek Limestone. The d13Corg isotope data collected was in-
tegrated with existing d13Corg data from this core (Sageman et al.,
In January 2012, Shell International Exploration and Production 2006; Du Vivier et al., 2014; Joo and Sageman, 2014).
Inc. drilled the research core Iona-1 (2913.510 N, 100 44.490 W),
located in the southern gateway of the Cretaceous Western Interior 2.4. Biostratigraphic methods
Seaway (KWIS; near the margin of the Gulf of Mexico carbonate
shelf in a distal, clastic-sediment-starved, intrashelf-sub-basin 2.4.1. Micropalaeontologic methods
setting along the retroarc foreland basin of western North America In total, 317 samples from Iona-1, Mustang-1 and USGS
(Fig. 1)). From the surface down, the core recovered the lower part Portland-1 were analyzed for quantitative micropaleontology. A
of the Austin Chalk (which crops out at this locality), a complete xed amount of sedimentary rock was gently crushed or dis-
record of the Eagle Ford Group and the upper part of the underlying aggregated (30 g dry weight), washed thoroughly through a 53 mm
Buda Limestone (Fig. 2). In total, 180 m of core spanning the earliest sieve and then oven-dried; normally a 63 mm is used, but a smaller
Cenomanian to earliest Coniacian was recovered. In the subsurface, sieve mesh was used to capture any calcispheres that might have
the Eagle Ford Group, characterized by organic-rich marlstones, otherwise been washed away. Depending on the quantity and
limestones, and numerous (>300) bentonites, can be divided into richness of the washed and dried residue a micro-splitter was used
two informal units, lower and upper, based on the transition from to subdivide the sample down to a size that contained at least 250
nely laminated organic-rich, marl-limestone couplets in the microfossils. The split factor was recorded for each sample along
lower to more bioturbated succession with more prominent with a dried weight to permit calculation of quantitative micro-
limestone beds in the upper Eagle Ford (Fig. 2; see Supplemental fossil abundance per gram. The split residue was then passed
Information Figs. S15eS16). Where the Eagle Ford Group crops through a set of sieves to provide four fractions: >500 mm,
out at the surface near this locality, it is termed the Boquillas For- 500e250 mm, 250e125 mm and <125 mm. The residue from each
mation, and has been informally divided into two members; the sieved fraction was spread on a picking tray and all microfossils
Langtry and Rock Pens Members (Pessagno, 1969). The underlying were hand-picked and deposited onto a slide for analysis. All
Buda Limestone and overlying Austin Chalk comprise massive specimens picked were quantitatively logged. The unpicked residue
limestones and chalks interbedded with thin marlstones and rare was then scanned for any further marker species; these were
bentonites. recorded in the database as occurring outside the count.
A

318
B D C E F

Depth (m)

marlstone
subsurface

limestone
Lithostrat.
Lithostrat.

bentonite
18Ocarb 13Ccarb 13Corg

Bentonite
outcrop
Grayscale

Isotope Event
beds
MTD
0 100 200 -10 -5 0 0 1 2 3 4 -28 -26 -24 -22
0

10
Austin Chalk
Austin Chalk

20

30

40

J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344


Langtry Mb.

50
Upper Eagle Ford

60

70

80
Boquillas Fm

D
Eagle Ford Gr

90 C IE-C

100 B IE-B
Rock Pens Mb.

A IE-A
110
Lower Eagle Ford

120

130 X

140
KWIS Bentonites

150
Limestone

160
Limestone

Buda
Buda

170

180

-10 -5 0 0 1 2 3 4 -28 -26 -24 -22


Fig. 2. Core data from Shell Iona-1. A. Lithostratigraphy: informal subsurface nomenclature in the Maverick Basin; outcrop lithostratigraphy (after Pessagno, 1969); lithologic log with bentonites, marlstone, limestone and mass
transport deposits (MTDs); stratigraphic distribution of bentonite beds with interpreted regional KWIS bentonites AeD, X after Elder, 1985, B. Grayscale values; C. Carbon isotope data from bulk carbonate (d13Ccarb); D. Oxygen isotope
data from bulk carbonate (d18Ocarb); E. Organic carbon isotope data (d13Corg) with isotopic events after (adapted from Pratt and Threlkeld, 1984); F. Identied global isotopic events (see text; Table 3); dark-gray horizontal
shading major isotopic events; light-gray horizontal shading earlier initiation of OAE-2 CIE based on inclusion of precursor events.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 319

Petrographic thin sections (10e30 mm thick) were counted via using standard crushing, heavy liquid, and magnetic separation
traverses across the samples until at least 300 conrmed specimens techniques, and were subsequently handpicked under the binoc-
were identied, and they were then scanned for any further marker ular microscope based on clarity and crystal morphology. When
species. grains were too small for standard crushing techniques they were
separated by using a shatterbox and clay separator (Hoke et al.,
2.4.2. Nannopaleontologic methods 2014). To overcome the effects of radioactive-decay-induced crys-
Nannofossil analyses were conducted on thickly concentrated tal defects and the associated lead-loss resulting in discordant an-
smear slides and involved counting all specimens within a standard alyses, zircon grains were pre-treated using a procedure modied
traverse (60 elds of view at 1000 magnication) plus the subse- after the thermal annealing and chemical abrasion isotope dilution
quent scanning of the remainder of the slide for rare species. In thermal ionization mass spectrometry method (CATIMs: of
richly nannofossiliferous samples, abundant species counts were Mattinson, 2005). This method involved heating of zircon inside a
estimated, with their counts from the rst traverses being multi- furnace at 900  C for 60 h. The annealed grains were subsequently
plied to equate with a full 60 eld of view traverse. loaded into FEP Teon microcapsules and leached in concentrated
HF at 210  C within high-pressure vessels for 12e18 h. The partially
2.4.3. Palynological methods dissolved sample was transferred into Savillex FEP beakers for
Sample splits were sequentially dissolved in hydrochloric (30% rinsing. The leached material was decanted with several milliliters
HCl) and hydrouoric (60% HF) acids. Post-HF oxidative treatment of ultra-pure water and by uxing successively with 4N HNO3 and
was necessary to liberate palynomorphs from prominent trans- 6N HCl on a hot plate and/or in an ultrasonic bath. After nal rinsing
lucent, uffy amorphous organic matter (AOM). Subsequent treat- with ultra-pure water, zircon grains were loaded back into their
ment with an alkali solution was necessary to remove post- microcapsules, spiked with a mixed 205Pbe233Ue235U tracer solu-
oxidation waste products. Previous work on C-torganic-rich tion (used in both laboratories to minimize inter-laboratory bias,
shales has indicated that excessive post-HF treatment led to se- made available to the community through the EARTHTIME project)
lective destruction of gonyaulacacean and areoligeracean dinocysts and dissolved completely in concentrated HF at 210  C for 48 h. U
(Dodsworth, 1995, 1996, 2004). Experiments with Eagle Ford post- and Pb were isolated from the zircon/tracer solutions using 50 mL
HF residues indicated that optimum oxidation was achieved with anion exchange columns (Krogh, 1973), the mixed, puried U and
extended maceration in cold nitric acid (HNO3). Samples containing Pb solutions were loaded onto rhenium laments and analyzed
prominent AOM were subjected to further maceration in warm with the mass spectrometer. The chemical abrasion method results
HNO3 (heated to 50  C) and treated with Schulze's Solution (70% in preferential dissolution of the high-U portions of the zircon
nitric acid supersaturated with potassium chlorate and/or fuming crystals that are more likely to have lost Pb leaving a residue with
nitric acid), followed by a single rinse in a detergent that contained relatively low U contents. This method has been found to be the
a trace of ammonia. This produces fairly uniform fossil assemblages best possible way to obtain the most concordant and highest pre-
with greatly reduced AOM concentrations, although scanning over cision analyses. Weighted mean 206Pb/238U ages from bentonites
many remaining AOM fragments was necessary to achieve suf- were used to anchor and constrain the astronomical age model and
cient specimen counts. Single Lycopodium tablets were added to rock accumulation rate calculations (post compaction).
samples after laboratory oxidation for absolute quantication of Sample ages are calculated based primarily on the weighted
palynomorphs (counts per gram; Mertens et al., 2009). Previous mean 206Pb/238U age of a statistically coherent cluster of analyses
experiments have shown that the oxidation required to liberate consisting of three or more data points using both ET_Redux
palynomorphs from AOM destroys a large proportion of Lycopo- (previously UePb_Redux; see Bowring et al., 2011) and Isoplot 3
dium spores if they are added prior to oxidation. Lund University, software (Ludwig, 2003), and is interpreted as the age of pyroclastic
Batch No. 1031 (2011) tablets were used. Each tablet contains eruption and a maximum estimate of the depositional age of the
20,848 Lycopodium spores (3457). As not all processed rock ma- bentonite. In a few bentonite samples where the zircon grain
terial was dissolved in HCl and HF, estimates of palynomorph population was dominated by xenocysts and a coherent cluster of
concentration are underestimates. However, values obtained are of analyses was not possible, the youngest zircon grain age is pre-
the same order as the only previously published KWIS study of the sented (see results and discussion sections). Uncertainties are re-
C-T palynomorph concentration (Dodsworth, 2000), i.e., they ported at 95% condence level and follow the notation X/Y/Z Ma,
generally range between 100 and 8000 specimens per gram. where X is the internal (analytical) uncertainty in the absence of all
Separate kerogen and post-oxidation biostratigraphy slides external errors, Y incorporates the UePb tracer calibration error,
were prepared. All preparations were sieved at 15 mm. The paly- and Z includes the latter as well as the decay constant errors of
nological counting technique used in this study involved making an Jaffey et al. (1971). Complete uncertainties (Z) are necessary for
initial count of 100 palynomorphs to estimate percentages of per- comparison between age data from different isotopic chronometers
idinioid, areoligeracean, ceratioid, Dinogymnium and gonyaulaca- (e.g., UePb versus 40Ar/39Ar). In this study, both UePb labs used the
cean dinocysts, Prasinophyte algae, acanthomorph acritarchs, same EARTHTIME tracer and thus can be compared using internal
foraminiferal test linings and terrigenous palynomorphs. Counting errors. All common Pb was assigned to laboratory blank.
of dinocysts was continued until a total of approximately 100
specimens were reached. Remaining slide material was thoroughly 2.6. Stable isotopic and geochemical methods
scanned for additional rare palynomorph taxa which were recorded
as outside the count in supplemental data tables. A total of 449 samples from Iona-1, Mustang-1 and USGS Portland
core were analyzed for bulk 13C/12C isotope ratios on the organic and
2.5. UePb geochronologic methods selected inorganic carbonate fractions (d13Corg and d13Ccarb); 18O/16O
isotope analyses were also performed on selected carbonate frac-
From the Iona-1 core, zircons from 11 bentonite beds (Fig. 2) tions. For the organic fraction, an aliquot of each powdered sample
were analyzed using the UePb ID-TIMS method at the Massachu- (0.1e2 g) was reacted with ~50 ml of 10% HCl overnight to remove
setts Institute of Technology (Cambridge, MA) and the University of any carbonate present, the acid was decanted, and then the sample
Toronto (Toronto, Canada) geochronology laboratories. Zircon and was rinsed three to four times with 250 ml of de-ionized water.
other U-bearing silicates were separated from bulk rock samples Carbonate fractions were analyzed using decanted acid aliquots. The
320 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

samples were dried in a 60  C oven and analyzed using a MAT253 All spectral analyses in this study for signal MTM used ve 3p
mass spectrometer at the Dept. of Earth Sciences Isotope Labs, discrete prolate spheroidal sequences (DPSS) tapers and those for
Southern Methodist University, Dallas, TX. Isotopic values are re- EHA and EPSA used three 2p DPSS tapers (Thomson, 1982). EHA and
ported relative to PDB; analytical uncertainties based on replicate EPSA were implemented using a 6-m moving window (0.01 m step),
and standard analyses are <0.1. Carbonate and organic C con- which was experimentally determined by using a range of possible
centrations were determined during the stable isotope analysis moving windows to be an optimal size to assess the change of the
workow. Total Organic Carbon (TOC) analyses were also performed spatial frequencies especially for tracing the strong obliquity signal.
on selected samples using a Leco C230 analyzer. Analytical un- However, due to the relatively constant frequency of the peaks
certainties for both methods are <0.1wt%. observed in the EHA spectrum from the lower Eagle Ford, indicative
of a stable rock accumulation rate, a larger 12-m MTM moving
2.7. Grayscale data window was applied to the grayscale data to obtain an EHA spectrum
with higher condence. Based on sensitivity analyses, the power
High-resolution (230 mm pixel) grayscale data were extracted spectra for windows with high signal to noise ratios centered at
from core-slab photographs of the Iona-1 research core (Fig. 2). The 106 m (100e112 m), 115 m (109e121 m), 120 m (114e126 m), and
grayscale extraction was averaged over a 5 mm wide longitudinal 125 m (119e131 m) were analyzed using ASM for determining the
transect and extracted using the software ImageJ. All core cracks orbital signals. For the upper Eagle Ford interval, the analyses were
and associated edge-effect reective features were removed from separated into several parts (39.32e50 m, 50e60 m, 60e75 m,
the dataset using MATLAB scripts and then manually validated. 75e85 m and 85e97.7 m) due to variable rock accumulation rates
Prior to analysis, data points associated with ~300 bentonites were determined from biostratigraphic and geochronologic constraints.
also removed from the time-series dataset. These deposits repre- MTM was then applied to these windows to determine the candidate
sent geologically instantaneous depositional events and generate peaks for ASM analyses. To obtain the signicant peaks, a sensitivity
both frequency and amplitude noise; the rapid addition of sedi- analysis was conducted to determine the threshold of F-test signif-
ment from volcanic activity distorts the frequency, and the light icance level ranging from >75% to >90%. Those signicant harmonic
color distorts the amplitude (Meyers et al., 2001). The resulting data components of the 12-m EHA window for the lower Eagle Ford and
were then depth adjusted for a continuous and uniformly sampled for different parts of the upper Eagle Ford served as the input for ASM
record and subjected to time-series analysis. Thus, all the depths analysis. The theoretical orbital target periods (henceforth the ETP
presented in astrochronologic methods and results sections are the model) for ASM analysis were derived from the astronomical model
adjusted depths, unless otherwise noted. This high-resolution of Laskar et al. (2004, 2011). Based on biostratigraphic and
grayscale data with Nyquist frequency of 4332 cycles/m enables geochronologic constraints, an orbital target for the upper Eagle Ford
the development of numerical astrochronologic age models. The was derived from Laskar solution in the interval 91e94.5 Ma span-
high-resolution grayscale was extracted from the Austin Chalk and ning the Cenomanian and Turonian, whereas the 94e98 Ma Laskar
the Eagle Ford Group intervals; the Buda Limestone was excluded solution was used for the lower Eagle Ford spanning the Cen-
from the analysis, due to pervasive irregular crack inllings and omanian. Following the approaches of Malinverno et al. (2010) and
mottled textures, as was the lowermost 6 m of the Eagle Ford Meyers et al. (2012a), mean values and 2s uncertainties of Laskar
Group, which is characterized by mass-transport deposits that solutions for seven Cretaceous astronomical periods were separately
were interpreted to represent geologically instantaneous deposi- obtained for the lower and upper Eagle Ford (see Supplementary
tional events. Data). The two dominant precession frequencies and two domi-
nant obliquity frequencies were evaluated via multitaper method
2.8. Astrochronologic methods (MTM harmonic analysis of eight 0.5 Ma segments of the Laskar et al.,
(2004) solution for lower Eagle Ford: 94e94.5, 94.5e95, 95e95.5,
Orbitally-forced signals preserved in the geological record are 95.5e96, 96e96.5, 96.5e97, 97e97.5, and 97.5e98 Ma and for the
complicated by the interference of random climate processes, seven segments for upper Eagle Ford: 91e91.5, 91.5e92, 92e92.5,
sediment accumulation rate changes, and a range of other 92.5e93, 93e93.5, 93.5e94, and 94e94.5 Ma. Using four eccen-
competing factors (Meyers et al., 2001, 2008; Weedon, 2003; Ma tricity solutions from Laskar et al. (2011), each eccentricity solution
et al., 2014). Consequently, more quantitative and statistical was evaluated in 2 Myr segments (94e96 and 96e98 Ma) for the
methodologies are required to constrain these noise sources and lower Eagle Ford and in segments of 91e92.5 and 92e94.5 Ma for the
explicitly test orbital hypotheses. In the present study, astronomical upper Eagle Ford yielding eight estimates of the two dominant short
signals were identied in the Iona-1 core by applying a combina- eccentricity terms. For the lower Eagle Ford and the upper Eagle
tion of the multi-taper method (MTM, Thomson, 1982), evolutive Ford, 4 Myr (94e98 Ma) and 3.5 Myr (91e94.5 Ma) segments
harmonic analysis (EHA), evolutive power spectral analysis (EPSA) respectively, were used to provide four estimates of the long
and average spectral mist (ASM, an inverse method; Meyers and eccentricity frequency (see Supplementary Data).
Hinnov, 2010; Meyers et al., 2012a; Ma et al., 2014) using the ASM hypothesis testing was conducted with 100,000 Monte
software astrochron: An R Package for Astrochronology (Meyers, Carlo simulated spectra, using the same number of frequencies as
2014). observed in each measured spectrum (see Meyers et al., 2012a).
The multi-taper method (MTM, Thomson, 1982) was used for One-hundred individual sediment accumulation rates were evalu-
identifying the signicant frequency associated with statistical ated, on a logarithmic grid, yielding a critical null hypothesis (H0)
signicance (F-test). Using the MTM method with a moving win- signicance level of 1%.
dow, EHA and EPSA were conducted to determine the change of the After determining the orbital signal, frequency bandpass was
spatial frequencies throughout the stratigraphic record. The ASM applied to extract the preserved orbital signals from the grayscale
was then quantitatively applied to t the spatial cycles with the record. Consequently, three astronomical age models were devel-
orbital targets to determine the optimal sediment accumulation oped using the following approaches: i) tracing the obliquity cycles;
rates as a function of time. Through the ASM method, Monte Carlo ii) the bandpass for long eccentricity associated with its stable
simulation was applied to reject the null hypothesis. Detailed de- period (Locklair and Sageman, 2008) and iii) ASM and were used to
scriptions of these methods can be found in Thomson (1982), calibrate the main biostratigraphic and isotopic events identied in
Meyers and Sageman (2007) and Ma et al. (2014). Iona-1 (Table 3).
Table 1
UePb analyses of zircons, Massachusetts Institute of Technology.
 206
Pbc 208
Pbd
Sample fractionsa Pb(c) (pg)b Pb b
Pbc
Th
U 204 Pb 206 Pb Ratios Date (Ma) Corr. coef.
206 207 207 206 207 207
Pbe Pbe Pbe Pb Pb Pb
238 U Error (2s%) 235 U Error (2s%) 206 Pb Error (2s%) 238 U Error (2s) 235 U Error (2s) 206 Pb Error (2s)

B1B; depth: 48.35e48.45 m


z1 0.22 79.7 0.53 4777 0.168 0.014201 (.06) 0.0936 (.30) 0.04783 (.29) 90.900 0.050 90.86 0.26 89.7 6.8 0.3
z4 0.18 51.1 0.52 3076.9 0.165 0.014189 (.06) 0.09341 (.43) 0.04777 (.40) 90.824 0.058 90.67 0.37 86.7 9.6 0.41
z5 0.18 27.5 0.61 1622 0.195 0.014172 (.08) 0.09349 (.93) 0.04787 (.90) 90.716 0.072 90.75 0.81 92 21 0.34
z6 0.26 23.1 0.60 1373.7 0.190 0.014179 (.09) 0.09319 (.93) 0.04769 (.90) 90.760 0.080 90.47 0.81 83 21 0.41
z7 0.21 15.1 0.71 877.3 0.226 0.014226 (.12) 0.09401 (1.28) 0.04795 (1.24) 91.06 0.11 91.2 1.1 96 29 0.36
z8 0.23 49.3 0.56 2935 0.178 0.014207 (.06) 0.0938 (.41) 0.04791 (.39) 90.942 0.053 91.04 0.36 94 9 0.38
Date (Ma) X Y Z
3 youngest analyses 90.776 0.039 0.060 0.11 MSWD 2.8
single youngest analysis 90.760 0.080 0.093 0.13

B2; depth: 81.5e81.6 m

z2 0.2 25.5 0.54 1537.3 0.172 0.014686 (.08) 0.09673 (.88) 0.04779 (.86) 93.985 0.073 93.75 0.79 88 20 0.37
z3 0.16 67 0.66 3880.6 0.209 0.018667 (.06) 0.12454 (.34) 0.04841 (.31) 119.228 0.070 119.18 0.38 118.2 7.4 0.46
z4 0.17 17.5 0.48 1074.5 0.154 0.014636 (.10) 0.09717 (1.12) 0.04817 (1.08) 93.669 0.094 94.2 1.0 107 25 0.44
z5 0.26 37.8 0.46 2314.1 0.146 0.014763 (.08) 0.09736 (.62) 0.04785 (.57) 94.470 0.078 94.34 0.56 91 14 0.59

J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344


z6 0.2 15 0.39 943.6 0.124 0.017225 (.14) 0.1184 (1.28) 0.04988 (1.23) 110.09 0.16 113.6 1.4 188 29 0.41
z8 0.2 57.9 0.14 3853.8 0.044 0.021243 (.06) 0.15512 (.31) 0.05298 (.29) 135.507 0.08 146.42 0.43 327.1 6.6 0.48
Date (Ma) X Y Z
single youngest analysis 93.669 0.094 0.11 0.15

B3; depth: 86.47 m

z4 0.2 90.3 0.56 5361.4 0.178 0.021476 (.06) 0.14437 (.29) 0.04878 (.27) 136.975 0.080 136.93 0.38 136.2 6.3 0.48
z3 0.3 22.3 0.73 1277.5 0.234 0.014706 (.09) 0.09693 (1.06) 0.04782 (1.04) 94.111 0.086 93.94 0.95 90 25 0.26
z2 0.3 34.8 0.69 2006.4 0.221 0.014642 (.06) 0.09686 (.56) 0.048 (.55) 93.705 0.057 93.88 0.5 98 13 0.18
z6 0.2 10.7 0.78 618.2 0.25 0.014636 (.16) 0.09604 (1.91) 0.04762 (1.84) 93.66 0.15 93.1 1.7 79 44 0.44
z1 0.2 22.4 0.68 1303.5 0.216 0.014634 (.09) 0.09663 (.96) 0.04791 (.92) 93.653 0.085 93.66 0.86 94 22 0.42
z5 0.2 38.6 0.8 2168 0.254 0.014629 (.07) 0.0968 (.52) 0.04801 (.50) 93.624 0.064 93.82 0.47 99 12 0.40
Date (Ma) X Y Z
4 youngest analyses 93.666 0.037 0.060 0.12 MSWD 1.2
single youngest analysis 93.624 0.064 0.079 0.13

B8; depth: 159.435 m

z1 0.19 9.8 0.84 560.0 0.267 0.015271 (.18) 0.10123 (2.12) 0.04810 (2.05) 97.70 0.18 97.9 2.0 103 48 0.40
z2 0.18 2.8 0.67 176.7 0.214 0.015252 (.55) 0.10003 (7.45) 0.04759 (7.25) 97.58 0.53 96.8 6.9 78 172 0.39
z4 0.17 29.8 0.75 1695.5 0.241 0.015255 (.08) 0.10060 (.94) 0.04785 (.92) 97.600 0.079 97.33 0.87 91 22 0.31
z5 0.14 32.8 1.10 1717.5 0.350 0.015292 (.11) 0.10136 (.84) 0.04810 (.80) 97.84 0.10 98.04 0.78 103 19 0.43
z6 0.19 22.4 0.74 1284.0 0.235 0.015251 (.12) 0.10097 (1.13) 0.04804 (1.07) 97.57 0.12 97.7 1.0 100 25 0.48
Date (Ma) X Y Z
3 youngest analyses 97.603 0.062 0.086 0.14 MSWD 0.73
single youngest analysis 97.57 0.12 0.13 0.17

B9; depth: 169.265 m

z1 0.4 12.6 0.56 767.1 0.177 0.015293 (.35) 0.10102 (1.37) 0.04793 (1.33) 97.84 0.12 97.7 1.3 95 32 0.33
z2 0.5 8.4 0.76 489.0 0.241 0.015301 (.17) 0.10144 (2.14) 0.04811 (2.09) 97.89 0.17 98.1 2.0 103 49 0.33
z3 0.4 18.6 0.80 1054.6 0.257 0.015296 (.54) 0.10118 (1.01) 0.04800 (.99) 97.855 0.088 97.86 0.94 98 23 0.30
z4 0.5 19.6 0.74 1123.6 0.237 0.015303 (.15) 0.10132 (.95) 0.04804 (.92) 97.901 0.095 97.99 0.89 100 22 0.35
z5 0.4 21.9 0.94 1194.9 0.300 0.015278 (.25) 0.10096 (.87) 0.04795 (.84) 97.745 0.091 97.66 0.81 96 20 0.36
Date (Ma) X Y Z
All 5 analyses 97.838 0.046 0.067 0.12 MSWD 1.6
single youngest analysis 97.745 0.091 0.10 0.15

321
322 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

3. Results
Xdinternal (analytical) uncertainty in the absence of all external or systematic errors; Ydincorporates the UePb tracer calibration error; Zdincludes X and Y, as well as the uranium decay constant errors. MSWDdmean square

Corrected for fractionation, spike, blank and initial Th/U disequilibrium in magma using Th/Umagma 2.8. All common Pb assumed to be blank. Mass fractionation correction of 0.25%/amu 0.04%/amu (atomic mass unit) was
3.1. Stable isotope results

Signicant carbon and oxygen isotopic differences are observed


both between and within marl and limestone lithofacies in all three
units (Austin Chalk, Eagle Ford and Buda Limestone), although the
All analyses are single zircon grains and pre-treated by the thermal annealing and acid leaching (CA-TIMS) technique. Data used in age calculations, as well as preferred calculated ages are in bold.

variability is most pronounced in the Eagle Ford. Some of this vari-


applied to single-collector Daly analyses. Total procedural blank less than 0.1 pg for U. Blank isotopic composition: 206Pb/204Pb 18.31 0.53, 207Pb/204Pb 15.38 0.35, 208Pb/204Pb 37.45 1.1.

ability can be attributed to early and late diagenetic processes


involving the formation of authigenic carbonate (e.g., Schrag et al.,
2013), and some can be attributed to variability in the isotopic
composition of carbon and oxygen in the water column from which
the biogenic carbonate that constitutes the matrix of these rocks was
derived. In ne-grained carbonate sediments, diagenetic alteration is
ubiquitous. Limestone-marl alternations, which often are interpreted
to reect cyclic changes in paleoenvironment, can also result from
differential diagenesis and differential compaction even where
paleoenvironmental cyclicity has been ruled out (Ricken, 1986;
Westphal et al., 2010). Aragonite and high-Mg calcite are particu-
larly susceptible to early dissolution with carbonate reprecipitated as
calcite cement. In the marine realm, calcite cementation can begin
on the seaoor in environments where, as was the case during Eagle
Ford deposition, active bacterial sulfate reduction produces elevated
porewater CO2 concentrations (Melim et al., 2002). In our dataset,
some limestones are characterized by relatively high d18Ocarb and
relatively low d13Ccarb values (Fig. 3) indicative of cements resulting
from CO2 production through shallow bacterial sulfate reduction
(Irwin et al., 1977) during early diagenesis (Fig. 3). However, we also
note that the range in values we report here are similar to those
recorded from pristine foraminifers from the KWIS (Fisher and
Arthur, 2002) and Tanzania (MacLeod et al., 2013) and thus re-
precipitated calcite may reect primary sea-water geochemistry.
Notes: Corr. coef. correlation coefcient. Age calculations are based on the decay constants of Jaffey et al. (1971).

Diagenetic complexities aside, bulk d13Ccarb data still represent ap-


proximations of the primary sea-water geochemistry as documented
by the global reproducibility and correlation of some of these d13Ccarb
trends that show minimal modication by diagenetic processes
(Immenhauser et al., 2008; Wendler, 2013). In particular, the Buda
Limestone recovered in Iona-1 is characterized by relatively uniform
d13Corg and d13Ccarb values (26 and 2 PDB, respectively), which
then decline in the basal Eagle Ford Group where they maintain
background levels of ~27.2 and 1e2, respectively. From these
background levels, ve notable excursions of varying magnitudes
were observed (Fig. 2; Table 3) including: i) a middle Cenomanian
Pbc is total common Pb in analysis. Pb* is radiogenic Pb concentration.

event (MCE; 143.73e139.27 m); ii) the C-T event encompassing OAE-
2 (105.96e91.56 m); iii) an early to middle Turonian event (EMTE;
78.05e68.87 m); iv) a middle Turonian event (MTE; 61.14e59.17 m);
and v) a late Turonian-Coniacian Event (LTCE; 49.44e5.85 m). In
Measured ratio corrected for spike and fractionation only.

particular, the C-T CIE is clearly expressed in our d13Corg record with a
positive CIE of up to 4 occurring between 105.96 and 92.73 m
(Fig. 2). However, our d13Corg isotopic record shows several precursor
isotope excursions of approximately ~1.5 between 111.36 and
107.17 m, which immediately precede and ultimately build up to the
C-T CIE. The shape and amplitude of the C-T CIE, as expressed in Iona-
1 is similar to other records in the KWIS (e.g. Bowman and Bralower,
2005; Sageman et al., 2006; Joo and Sageman, 2014) with the intra
CIE peaks as identied by Pratt and Threlkeld (1984) being clearly
resolved (Fig. 2). The d13Ccarb data exhibit similar trends, although
they are of lower amplitudes (1e2).
Radiogenic Pb ratio.
of weighted deviates.

3.2. Biostratigraphical results

Recovery of microfossils from all three disciplines (micropa-


leontology, nannopaleontology, palynology) throughout the
Eagle Ford and Austin Chalk intervals was excellent with the
a

e
b
c
d

identication of numerous age diagnostic bioevents (Fig. 4;


J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 323

Table 2
UePb analyses of zircons; University of Toronto.
 206
Pbc 208
Pbd
Sample Pb(c) Pb c Th Ratios Date (Ma) Corr.
Pbc U 204 Pb 206 Pb

fractionsa (pg)b 206


Pbe 207
Pbe 207
Pbe 206
Pb 207
Pb 207
Pb coef.
238 U Error 235 U Error 206 Pb Error 238 U Error 235 U Error 206 Pb Error
(2s%) (2s%) (2s%) (2s) (2s) (2s)

B1A; depth: 48.06 m


z1 1.9 21 1.91 511 0.6062 0.017303 (.30) 0.1206 (.35) 0.0505 (.14) 110.59 0.3 115.6 3.1 220 63 0.683
z2 1.2 26 1.25 1133 0.3973 0.016303 (.13) 0.1107 (.13) 0.0493 (5.35) 104.25 0.13 106.6 1.2 160 25 0.662
z3 0.9 17 0.9 1106 0.2853 0.015232 (.16) 0.1136 (.12) 0.0541 (5.43) 97.45 0.16 109.2 1.1 374 23 0.573
z4 1.3 16 1.34 615 0.4262 0.014971 (.22) 0.0990 (.22) 0.0480 (9.93) 95.79 0.22 95.9 2 98 49 0.623
z5 0.9 16 0.94 1026 0.2971 0.014701 (.12) 0.0953 (.13) 0.0470 (5.97) 94.08 0.12 92.4 1.2 50 30 0.683
z6 1 21 0.69 1286 0.2193 0.014516 (.14) 0.0954 (.17) 0.0477 (8.09) 92.90 0.14 92.6 1.6 84 40 0.728
Date (Ma) X Y Z
single youngest analysis 92.90 0.14 0.15 0.18

B4; depth: 100.66 m

Z1 0.3 2.4 0.76 462 0.2399 0.014826 (.51) 0.0999 (.91) 0.0488 (.42) 94.87 0.51 96.6 8.4 141 197 0.967
z2 0.3 3.7 0.59 760 0.1859 0.014784 (.21) 0.0984 (.27) 0.0483 (.12) 94.61 0.21 95.3 2.5 112 61 0.753
z3 0.3 6.3 0.64 1203 0.2031 0.014741 (.15) 0.0970 (.19) 0.0477 (.09) 94.33 0.15 94 1.7 86 43 0.743
Date (Ma) X Y Z
All analyses 94.45 0.49 0.51 0.53 MSWD 3.8
single youngest analysis 94.33 0.15 0.17 0.20

B5; depth: 114.04 m

z1 2.3 7 0.4 209 0.1254 0.019744 (.59) 0.1362 (.93) 0.0500 (.32) 126.04 0.59 129.7 8.4 197 153 0.867
z2 2.6 11.7 1.2 253 0.3812 0.016962 (.45) 0.1159 (.63) 0.0496 (.25) 108.43 0.45 111.4 5.8 175 122 0.79
z3 3.7 11.8 0.89 199 0.2814 0.016428 (.44) 0.1118 (.80) 0.0494 (.33) 105.04 0.44 107.6 7.3 165 161 0.981
z4 2 8.3 1.96 204 0.6234 0.016406 (.49) 0.1094 (.80) 0.0483 (.33) 104.9 0.49 105.4 7.3 116 166 0.894
z5 1.8 7.7 2.19 203 0.6972 0.016247 (.44) 0.1080 (.77) 0.0482 (.33) 103.89 0.44 104.1 7.1 109 163 0.955
z6 1.4 9.1 0.53 417 0.1670 0.015387 (.22) 0.1024 (.35) 0.0483 (.15) 98.44 0.22 99 3.2 113 76 0.881
Date (Ma) X Y Z
single youngest analysis 98.44 0.22 0.23 0.25

B6; depth: 124.78 m

z1 0.4 6.6 0.44 944 0.1408 0.015042 (.19) 0.1022 (.20) 0.0493 (.09) 96.25 0.19 98.8 1.9 160 44 0.663
z2 0.4 6.1 0.62 973 0.1956 0.014985 (.17) 0.0982 (.16) 0.0475 (.07) 95.88 0.17 95.1 1.4 76 35 0.596
z3 1.6 8.3 0.43 337 0.014959 (.26) 0.0988 (.42) 0.0479 (.19) 95.72 0.26 95.7 3.9 95 97 0.907
z4 0.4 9.4 0.48 1417 0.1535 0.014934 (.18) 0.0995 (.13) 0.0483 (.06) 95.56 0.18 96.3 1.2 115 28 0.525
z5 0.5 10 0.45 1265 0.1434 0.014921 (.12) 0.0992 (.12) 0.0482 (.05) 95.48 0.12 96 1.1 110 26 0.64
z6 0.4 6.2 0.37 898 0.1179 0.014894 (.17) 0.0989 (.19) 0.0482 (.09) 95.30 0.17 95.8 1.8 108 43 0.695
Date (Ma) X Y Z
4 youngest analyses 95.47 0.22 0.23 0.24 MSWD 2.9
single youngest analysis 95.30 0.17 0.18 0.20

B7; depth: 127.12 m

z1 0.4 10 0.12 1911 0.1782 0.015207 (.13) 0.1038 (.08) 0.0495 (.03) 97.29 0.13 100.3 0.7 172 16 0.539
z2 0.6 11.8 0.51 1238 0.1612 0.014983 (.16) 0.0996 (.12) 0.0482 (.05) 95.87 0.16 96.4 1.1 110 26 0.54
z3 0.7 13 0.56 1217 0.1782 0.014968 (.13) 0.0981 (.12) 0.0475 (.06) 95.78 0.13 95 1.1 75 28 0.626
z4 1.3 13.8 0.5 649 0.1598 0.01496 (.28) 0.098 (.27) 0.0475 (.12) 95.73 0.28 94.9 2.5 75 61 0.593
z5 0.5 7.1 0.53 870 0.1679 0.014952 (.46) 0.0987 (.16) 0.0479 (.07) 95.68 0.46 95.6 1.5 94 35 0.464
z6 0.8 15.2 0.46 1238 0.1450 0.014793 (.26) 0.0981 (.11) 0.0481 (.05) 94.66 0.26 95 1.1 104 25 0.47
Date (Ma) X Y Z
4 youngest analyses (excl. z6) 95.80 0.092 0.11 0.14 MSWD 0.48
single youngest analysis (Pb 94.66 0.26 0.29 0.31
loss?)

Notes: Corr. coef. correlation coefcient of XeY error on the Concordia plot. Age calculations are based on the decay constants of Jaffey et al. (1971).
Xdinternal (analytical) uncertainty in the absence of all external or systematic errors; Ydincorporates the UePb tracer calibration error; Zdincludes X and Y, as well as the
uranium decay constant errors. MSWDdmean square of weighted deviates.
a
All analyses are single zircon grains and pre-treated by the thermal annealing and acid leaching (CA-TIMS) technique. Data used in age calculations, as well as preferred
calculated ages, are in bold.
b
Pbc is total amount of common Pb in analysis. Pb* is radiogenic Pb concentration. Th/U calculated from radiogenic 208Pb/206Pb ratio and 207Pb/206Pb date assuming
concordance.
c
Measured ratio corrected for common Pb in spike and fractionation.
d
Radiogenic Pb ratio.
e
Corrected for fractionation, spike, blank and initial Th/U disequilibrium in magma using Th/Umagma 4.2. All common Pb assumed to be blank.

Table 3; see Supplementary Data). However, recovery of mi- is the authors intention to publish detailed taxonomic infor-
crofossils from the Buda Limestone was moderate due to the mation elsewhere.
more indurated limestones limiting the extraction of forami-
nifera, although several age diagnostic events were identied. 3.3. Bentonite UePb geochronologic results
The raw data and detailed biostratigraphic discussion of the
bioevents for nannofossils, foraminifera and dinoagellate cysts The range in zircon dates reects complex populations indi-
are presented in the supplementary information sections, and it cating a variety of processes, including the incorporation of
324 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Table 3
List of calibrated bio-isotopic events for the Cenomanian-Turonian of the KWIS. The following nomenclature was used for the denition of bio-events: LO last occurrence;
LCO last common occurrence; FO rst occurrence, FCO rst common occurrence; cons. consistent occurrence (see supplemental data for more details). Code letter
follows microfossil discipline and number is stratigraphically oldest to youngest. Selected taxa are presented in the Plates 1e4.

Event Code Depth (m) O1 tracing (Ma) (2s) E1 bandpass (Ma) (2s) ASM and UePb interpolation (Ma) (2s)

Planktonic Foraminifera
LO Favusella washitensis F1 171.51 e e e e 97.90 0.12
LO Rotalipora brotzeni F2 145.75 96.67 0.13 96.73 0.17 e e
LO Thalmanninella appenninica F3 143.37 96.56 0.13 96.62 0.17 e e
FO Whiteinella aprica F4 140.39 96.42 0.12 96.50 0.16 e e
FO Helvetoglobotruncana praehelvetica F5 136.88 96.23 0.12 96.35 0.16 e e
FO Rotalipora cushmani F6 133.68 96.06 0.13 96.20 0.17 e e
FO Whiteinella archaeocretacea F7 130.74 95.92 0.15 96.07 0.18 e e
FO cons. Rotalipora cushmani F8 116.76 95.23 0.12 95.30 0.17 e e
FO cons. Whiteinella archaeocretacea F9 116.76 95.23 0.12 95.30 0.17 e e
LO Rotalipora deeckei F10 106.94 94.73 0.12 94.75 0.16 e e
LO Thalmanninella greenhornensis F11 105.96 94.67 0.12 94.69 0.16 e e
LO Rotalipora cushmani F12 105.40 94.65 0.12 94.68 0.16 e e
FO Dicarinella hagni F13 104.83 94.62 0.12 94.66 0.16 e e
FO Guembelitria cenomana F14 104.83 94.62 0.12 94.66 0.16 e e
FAO Heterohelix reussi-moremani (Heterohelix shift) F15 100.90 94.43 0.13 94.54 0.17 e e
FO Marginotruncana renzi F16a 87.88 93.76 0.13 93.94 0.21 e e
FO Marginotruncana schneegansi F16b 87.88 93.76 0.13 93.94 0.21 e e
FO Helvetoglobotruncana helvetica F17 79.51 93.21 0.13 93.13 0.17 e e
FO Marginotruncana coronata F18 64.50 92.24 0.13 91.90 0.17 e e
LO Praeglobotruncana gibba F19 56.61 91.49 0.15 91.28 0.17 e e

Calcareous Nannoplankton

FO Corollithion kennedyi N1 136.88 96.23 0.12 96.35 0.16 e e


ACME Corollithion kennedyi N2 112.77 95.02 0.13 95.07 0.17 e e
FO Gartnerago obliquum N3 112.77 95.02 0.13 95.07 0.17 e e
LO Axopodorhabdus albianus N4 107.17 94.74 0.12 94.76 0.16 e e
LO Corollithion kennedyi N5 104.18 94.59 0.12 94.64 0.16 e e
LO Cretarhabdus striatus N6 104.18 94.59 0.12 94.64 0.16 e e
LO Helenea chiastia (Iona-1) N7a 101.83 94.43 0.12 94.54 0.16 e e
LO Helenea chiastia (extrapolated) N7b ~94.50 94.10 0.12 94.31 0.16 e e
FO Quadrum gartneri N8 90.57 93.90 0.12 94.15 0.16 e e
FO Eprolithus octopetalus N9 89.60 93.85 0.17 94.11 0.18 e e
FO ?Eprolithus eptapetalus N10 87.88 93.76 0.13 93.94 0.21 e e
LO Eprolithus octopetalus N11 81.40 93.33 0.14 93.26 0.18 e e
FO Eiffellithus eximius N12 77.60 93.11 0.12 92.99 0.17 e e
FCO Eiffellithus eximius N13 65.80 92.32 0.13 92.00 0.17 e e
LO Radiolithus orbiculatus N14 58.44 91.68 0.17 91.42 0.19 e e
FO Lithastrinus septenarius N15 56.61 91.49 0.13 91.28 0.17 e e
FO Lucianorhabdus maleformis N16 40.35 e e e e 90.23 e
FO Marthasterites furcatus N17 18.67 e e e e 89.72 e
LO Stoverius achylosus N18 12.84 e e e e 89.52 e

Radiolarians

FAO Radiolarians R1 103.82 94.57 0.12 94.63 0.16 e e

Dinoagellate Cysts

FCO Litosphaeridium siphoniphorum P1 144.21 96.60 0.12 96.66 0.16 e e


LSAO Bosedinia spp. cf. sp.1 & sp.3 of Prauss (2012b) P2 120.40 95.41 0.13 95.51 0.17 e e
LCO Cyclonephelium longispinatum P3 119.49 95.36 0.13 95.46 0.17 e e
FO Cyclonephelium membraniphorum P4 105.96 94.68 0.12 94.69 0.17 e e
LO Adnatosphaeridium tutulosum P5 104.18 94.59 0.13 94.64 0.16 e e
LO cons. Litosphaeridium siphoniphorum P6 103.44 94.56 0.98 94.62 0.98 e e
LO Adnatosphaeridium? chonetum P7 88.77 93.81 0.13 94.08 0.21 e e
FO Eurydinium glomeratum P8 87.88 93.76 0.13 93.94 0.21 e e
FO Chatangiella spectabilis P9 74.50 92.95 0.13 92.77 0.17 e e
LCO Cyclonephelium membraniphorum P10 72.77 92.87 0.12 92.66 0.16 e e
FO Heterosphaeridium difcile P11 65.12 92.28 0.13 91.95 0.17 e e
FO cons. Senoniasphaera turonica P12 60.93 91.94 0.12 91.61 0.16 e e
FCO Senoniasphaer rotundata sensu stricto P13 37.89 e e e e 90.17 e
LO Isabelidinium magnum P14 7.91 e e e e 89.45 e

d13Corg Carbon Isotope Events

Middle Cenomanian Event (MCE) start MCEs 143.73 96.57 0.12 96.64 0.16 e e
Middle Cenomanian Event (MCE) end MCEe 139.27 96.36 0.13 96.45 0.17 e e
C-T Precursor Event 1 111.36 94.94 0.13 94.99 0.17 e e
C-T Precursor Event 2 108.77 94.81 0.12 94.85 0.16 e e
C-T Precursor Event 3 107.17 94.74 0.12 94.76 0.16 e e
C-T CIE Event A (Pratt & Threlkeld, 1984) IE A 105.40 94.65 0.12 94.68 0.16 e e
C-T CIE Event B (Pratt & Threlkeld, 1984) IE B 100.90 94.43 0.12 94.54 0.16 e e
C-T CIE Event C (Pratt & Threlkeld, 1984) IE C 92.73 94.02 0.12 94.24 0.16 e e
Early-Middle Turonian Event start EMTEs 78.05 93.13 0.12 93.02 0.16 e e
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 325

Table 3 (continued )

Event Code Depth (m) O1 tracing (Ma) (2s) E1 bandpass (Ma) (2s) ASM and UePb interpolation (Ma) (2s)

EarlyeMiddle Turonian Event end EMTEe 66.49 92.56 0.13 92.22 0.17 e e
Middle Turonian Event start MTEs 61.14 91.95 0.12 91.63 0.16 e e
Middle Turonian Event end MTEe 59.17 91.76 0.13 91.47 0.17 e e
Late Turonian-Coniacian Event start LTCEs 49.44 90.83 0.13 90.75 0.17 e e
Late Turonian-Coniacian Event end LTCEe 43.78 90.42 e e e 90.05 e

A 6 B 0
4

organic matter oxidation


Inoceramus

meteroic
water
marine
-2
13Ccarb ()

vug cements
2

18Ocarb ()
pelagic
ooze
0 -4
recent
-2 sediments

-4 -6
Inoceramus

-6 meteroic water -8
vug
-8 concretions
-10 -10
-10 -8 -6 -4 -2 0 2 4 6 0 20 40 60 80 100
18Ocarb () Calcite (wt. %)

C 4 D 15
Inoceramus

2 12
TOC (wt. %)
13Ccarb ()

vug

0 9
organic matter
oxidation

-2 6

-4 3

-6 0
0 20 40 60 80 100 0 20 40 60 80 100
Calcite (wt. %) Calcite (wt. %)
E 320 F 320
280 280
2
240 240 r = 0.51
Grayscale
Grayscale

200 200
160 160
120 120
80 2 80
r = 0.59
40 40
0 0
0 20 40 60 80 100 0 3 6 9 12 15
Calcite (wt. %) TOC (wt. %)
Fig. 3. Cross plots showing: A. relationship between d13C and d18O to elucidate the primary vs diagenetic signal; oval elds from Moore (1989); dashed box range of pristine
Cenomanian-Turonian Foraminifera recovered from Tanzania (MacLeod et al., 2013); solid box pristine Cenomanian-Turonian foraminifera from the KWIS (Fisher and Arthur,
2002). B. d18Ocarb and calcite (wt %); C. d13Ccarb and calcite (wt. %); D. TOC (wt. %) and calcite (wt. %); E and F. correlation between grayscale with calcite (wt %) and TOC (wt
%), respectively.
326 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

A B C D E F

Bentonite Beds

Isotope Event
Radiolarian
13Corg
Depth (m)
subsurface

Geochron.
marlstone
limestone

bentonites
bentonite

Dinocysts
Lithostrat.

Lithostrat.

Regional

U/Pb
outcrop

Nannofossils Foraminifera

MTD
-28 -26 -24 -22

P14
Austin Chalk

Austin Chalk

10
N18
N17
20
UC9

M. schnegansi
30

P13
40 N16

B1A
LTCE
Langtry Mb.

50 B1B
N15 F19
N14 P12 MTE
60
Upper Eagle Ford

UC8 N13 F18 P11


H. helvetica

70 P10 EMTE
P9
N12 F17
80 B2 N11
UC7
D B3 N10 F16 P8
Boquillas Fm

90 N9 P7
C N8 IE-C
W. arch.

UC6
Rock Pens Mb.

N7b OAE-2
100 B B4 UC5
N7a
F14 F15 P6
IE-B
A F13 P5 R1
N5 N6 F12 P4
IE-A
N4 F11 precursor events
110 F10
B5 N2
N3 F8
Lower Eagle Ford

UC1 - UC4

R. cushmani

120 F9 P3
P2
B6
X B7
130 F7
F6
N1 F5
140 F4 MCE
F3
F2
150
not assigned
Limestone

Limestone

B8 P1
160
Buda

Buda

170 B9
F1

180

-28 -26 -24 -22


Fig. 4. Summary of biostratigraphic and chemostratigraphic events in the Iona-1 core. A. Lithostratigraphy and lithology (see Fig. 2 for legend); stratigraphic distribution of
bentonite beds; B. Regional KWIS bentonites AeD, and X of Elder (1985); C. Selected bentonites used for UePb geochronology; D. Selected biostratigraphic marker events (Table 3;
Supplemental Information) and biozones (UC zonation after Burnett, 1998; foraminiferal zonation after ); E. Organic carbon isotope data (d13Corg) with isotopic events after (adapted
from Pratt and Threlkeld, 1984); F. Identied global isotopic events (see text; Table 3).
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 327

inherited components in the zircon population, recycling of some youngest zircon grain in bentonites B1A (48.06 m) and B5
zircon grains that formed during protracted magmatic evolution of (114.04 m) are still much older than those derived from other
a magmatic complex and interaction of the source magmas with stratigraphically proximal bentonites as well as the astronomical
their country rocks, and/or post-eruption mixing of zircons in the age model and are interpreted as xenocrysts; therefore, they do not
sedimentary environment (Fig. 5). In addition, the selection of reect the eruptive/depositional age of the bentonites (Fig. 5).
which zircon dates provides the best estimate of the eruptive/
depositional age requires an assessment of whether chemical 3.4. Astronomical results
abrasion eliminated the effects of Pb loss and which zircons are
older than the eruption age. In general, we selected the best- Extracted grayscale values provide a reasonable proxy for the
formed, inclusion-free, and clear grains prior to analysis. Once geochemistry of the two main lithofacies alternating in the
analyzed, we eliminated obvious outliers due to inheritance or analyzed stratigraphic sections (dark marlstones and light lime-
Pb-loss and then evaluated the remaining grains. The approach of stones). Changes in grayscale values positively and negatively
calculating a weighted mean from multiple analyses of the correlate with the quantity of carbonate and TOC (wt %), respec-
remaining grains assumes that they represent a single-aged pop- tively, as shown in Fig. 3E and F. Some variation is expected as the
ulation and that scatter or dispersion in the data is the result of geochemical measurements represent the average composition of
analytical uncertainty. A mean standard weighted distribution many laminae as they are derived from bulk 2.5 cm diameter core
(MSWD) of approximately 1 is consistent with this assumption and plug samples from the 2/3 core portion, whereas the grayscale,
the weighted mean is taken to be the best estimate of the crystal- although averaged over a 3-cm interval, represents the surface
lization age of the zircon and in this case the eruption/depositional color value of the slabbed core at slightly different lateral locations
age as well (Tables 1 and 2). Due to the high closure temperature for in the 1/3 core slab portion, but at similar depths.
UePb in zircon (>900  C for moderate-sized grains; Cherniak and The cyclostratigraphic study focuses on spatial cycles that fall
Watson, 2003) and rapid cooling rates, individual zircon grains within the orbital bands, as determined by the nominal zircon
that are xenocrystic with respect to the magmatic systems may UePb age constraints (Tables 1 and 2) and supported by the age
reect processes that predate the eruption and/or inclusion of model. Considering the sediment accumulation rate constraint,
basement or country-rock material incorporated in the ash cloud frequencies of 6 cycles/m and 10 cycles/m are used as the largest
during the explosive Plinian eruptions. Younger UePb ages could be spatial frequencies tested in ASM analyses for the lower and upper
the result of residual post-crystallization Pb loss; since the devel- Eagle Ford, respectively (Ma et al., 2014). The sediment accumula-
opment of the CA-ID-TIMs method (Mattinson, 2005), the occur- tion rates tested in ASM are 1e3 cm/kyr and 0.4e6 cm/kyr for the
rence of UePb dates that record Pb loss have been greatly reduced. lower and upper Eagle Ford, respectively, based on the zircon UePb
Within our dataset (Fig. 5, Tables 1 and 2), only the youngest grain age constraints (Tables 1 and 2). The Average Spectral Mist (ASM)
from bentonite B7 (z6; 127.12 m) may have experienced minor Pb results are presented in Table 4 and also in the supplemental
loss and was excluded from the age calculation. Furthermore, the information.

Age B9M B8M B7T B6T B5T B4T B3M B2M B1BM B1AT
(Ma) (169.265m) (159.435m) (127.12m) (124.78m) (114.04m) (100.66m) (86.47m) (81.5m) (48.35m) (48.06m)
90.0
z8 z5 z1
z4 z6 z7
91.0

92.0

z6
93.0
z2 z6 z1 z5 z4
z3 z2 z5
94.0 z3
z6 z1 z2 z5

95.0 z6
z5 z3 z4 z5
z4
z4
z2 z3 z2
96.0 z1

z2
97.0 z1 z3
z1 z4 z6
z5 z5
z4 z2 z3 z1
98.0 z6

99.0 z1 z2 z3 z4 z5 z8 z3 z6 z1 z2
bar heights are 2
100.0

Fig. 5. Age distribution plots of UePb ID-TIMS analyses. Bentonites with superscript T analyzed at the University of Toronto; those with M at the Massachusetts Institute of
Technology. Bar heights show 2-sigma internal uncertainties of individual zircon analyses; dark-gray bars those used in age calculation; light-gray bars those excluded from age
calculation (see text for discussion). Older analyses plotting outside the diagrams are marked with arrows. Shaded horizontal line is the proposed age of the C-T boundary. See
Tables 1 and 2 for the complete analytical data, calculated dates, and their uncertainties.
328 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Table 4 respectively. For the interval of 85e97.7 m (adjusted depth), long


Average rock accumulation rates obtained from ASM results for Eagle Ford Gr. eccentricity (E1), short eccentricity (E2, E3), obliquity (O1, O2), and
Interval (m) Rock accumulation rate (cm/ka) precession (P1, P2) were bandpassed at frequency intervals of
39.32e50 1.01
0.05e0.15, 0.3e0.5, 0.75e1.25, and 1.85e2.4 cycles/m, respectively.
50e60 0.99 Based on the bandpass of long eccentricity with the relatively stable
60e75 1.80 period of 405.16 (lower Eagle Ford) and 402.41 ka (upper Eagle
75e85 1.41 Ford), astronomical age model 2 was established by using the long
85e97.7 1.78
eccentricity bandpass. The continuity of the long eccentricity
97.7e110.5 1.54
110.5e117.5 1.95 bandpass signal was checked by performing the analyses from both
117.5e122.5 1.70 bottom-up and top-down and provided similar results.
122.5e133.652 1.67
3.4.1.3. Astronomical age model 3 (ASM). Astronomical age model 3
was constructed using the optimal rock accumulation rates ob-
tained from ASM results (Fig. 7). This age model was only used for
3.4.1. Astronomical age model results the Austin Chalk interval, where the rock accumulation rates were
Three alternative age models were developed based on: i) steady (~3.87 cm/ka) and O1 tracing was not conducted.
orbital tracing of the O1 signal; ii) ASM; as well as iii) E1 eccen-
tricity bandpass (see Methods), and they were used to calibrate the 4. Astronomical age calibration
main biostratigraphic and isotopic events identied in Iona-1 as
detailed in supplemental information. The E1 long eccentricity The development of three astronomical age models for Iona-1
term has been predicted to be the most stable and phase-coherent has enabled the direct age calibration of the main age-diagnostic
astronomical term, providing a superior tuning target (Laskar et al., micropaleontological datums and global stable isotope curve for
2004, 2011); however, the preserved sedimentary record of long the Cenomanian and Turonian stages in the KWIS. This calibration
eccentricity is particularly sensitive to stochastic noise which can has the potential to be applied to other global localities to provide
distort the reconstructed bandpass-lter output (see Meyers, 2012; absolute age constraints. The main biostratigraphic and isotopic
Ma et al., 2014). Alternatively, the O1 astronomical term is less events calibrated against the astronomical age models are pre-
stable and difcult to predict (Laskar et al., 2004, 2011), but the sented in Table 3 and are discussed in the supplemental informa-
preserved O1 signal in Iona-1 is persistent and shows a relatively tion. Time-series analyses were not performed for the Buda
strong high amplitude bedding signal throughout the studied sec- Limestone, so the assigned ages are based on a linear regression
tion providing a higher resolution age model. Both the E1 eccen- between dated bentonite layers. In addition, the assigned age for
tricity bandpass and O1 tracing astronomical models are presented the lowermost 6 m of the Eagle Ford Group, which contain mass-
for the Eagle Ford section and used to calculate the age of the C-T transport deposits, was calculated by linear extrapolation of the
boundary (see discussion), whereas the ASM model was only used overlying trend. To constrain the stratigraphic occurrence of some
for the overlying Austin Chalk section. The E1 bandpass age model biostratigraphic and isotopic events, we compared the Iona-1 core
provides higher tuning condence, but provides lower resolution data with those from a nearby core, Mustang-1, and with the
than the O1 tracing age model. astrochronologically tuned USGS Porltand-1 core, near the C-T
GSSP, Pueblo (Meyers et al., 2001, 2012a; Sageman et al., 2006;
Fig. 9). The age-calibration approach follows that of Sageman et al.
3.4.1.1. Astronomical age model 1 (orbital O1 tracing). Based on the
(2014) and includes total propagated uncertainties at 2s based on:
sediment accumulation rates from the above ASM analyses, the
i) analytical and total uncertainty of the geochronological mea-
orbital targets were recognized and labeled on the EHA and EPSA
surements at 95% condence level, ii) depth uncertainty of the bio-
results in Fig. 6. Here the strong signals of obliquity (O1), with a
and isotopic events due to the sampling resolution, and iii) uncer-
period of 50.23 ka for the lower Eagle Ford and 49.82 ka for the
tainty in the astrochronologic methods. Furthermore, for the cali-
upper Eagle Ford were traced and the corresponding spatial fre-
bration of the C-T stage boundary, geological uncertainty related to
quencies were used for constructing the astronomical age model 1
the stratigraphic correlation between the GSSP, Pueblo and the
(Fig. 7).
Iona-1 core is also included (see supplemental information for
details). In addition, following the methods of Sageman et al.
3.4.1.2. Astronomical age model 2 (Eccentricity). For each preserved (2014), the C-T stage boundary age estimate was also calculated
orbital target, the bandpass signal was extracted from the grayscale using multiple isotopically dated bentonite layers to anchor both
data (Fig. 8). For the lower Eagle Ford, long eccentricity (E1), short the O1 tracing and E1 bandpass astronomical age models (see
eccentricity (E2, E3), obliquity (O1, O2), and precession (P1, P2) Discussion).
were bandpassed at the frequencies of 0.08e0.16, 0.4e0.63, 0.9e1.7,
and 2.1e3.3 cycles/m respectively. As the rock accumulation rates 5. Discussion
were similar for the intervals 39.32e50 and 60e50 m (Fig. 6); the
206
bandpasses were conducted together. Long eccentricity (E1), short 5.1. Comparison of Pb/238U with 40
Ar/39Ar dating
eccentricity (E2, E3), obliquity (O1, O2) and precession (P1, P2) were
bandpassed at frequency intervals of 0.15e0.23, 0.7e1.2, 1.9e2.8, Signicant advances in both the 206Pb/238U and 40Ar/39Ar iso-
and 4.0e5.8 cycles/m, respectively. For the interval of 60e75 m topic methodology including improved standardization and inter-
(adjusted depth), long eccentricity (E1), short eccentricity (E2, E3), laboratory calibration (e.g., Mattinson, 2005, 2010; Condon et al.,
obliquity (O1, O2), and precession (P1, P2) were bandpassed at 2007; Kuiper et al., 2008; Renne et al., 2010; Schmitz and
frequency intervals of 0.15e0.23, 0.4e0.65, 0.9e1.5, and 2.3e3.1 Bowring, 2001) now enable higher precision dating of both apa-
cycles/m, respectively. For the interval of 75e85 m (adjusted tites and zirons for geochronologic applications. Each method has
depth), long eccentricity (E1), short eccentricity (E2, E3), obliquity advantages and disadvantages which is beyond the scope of this
(O1, O2), and precession (P1, P2) were bandpassed at frequency paper. We refer the interested reader to the detailed discussions in
intervals of 0.2e0.3, 0.55e0.75, 1.3e2.2, and 3.0e3.9 cycles/m, Renne (2014) and Sageman et al. (2014).
A Grayscale B C D E Rock accumulation
0 100 200
EHA - Normalized and Filtered Amplitude Rate (cm/ka)
EHA results (Amplitude) EPSA results (Log Power)
40 40
(Black <80% Harmonic CL)

O2 1.0 O2 O2
E2+3 E2+3 E2+3
50 P1 P2 P1 P2 P1 P2 4 50
E1 O1 E1 O1 25 E1 O1

60 60
0.8
E1 E1 E1

J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344


O2 P1 P2 O2 P1 P2 20 O2 P1 P2
E2+3 E2+3 E2+3 2
70 70
Adjusted Depth (m)

O1 O1 O1

Adjusted Depth (m)


E1 0.6 E1 E1
80 80
O2 P1 P2 O2 P1 P2 15 O2 P1 P2

0
90 90

P1 P2 0.4 P1 P2 P1 P2
E1 E1 10 E1
100 100

O1 P1 P2 O1 P1 P2 O1 P1 P2
E1 E1 E1 2
110 E2+3 E2+3 E2+3
O2 O2 O2 110
0.2 5

E1 E1 E1
120 120

4
P1 P2 P1 P2 P1 P2
130 130
133.32 0 1 2 3 4 5 6 0 1 2 3 4 5 6 0 1 2 3 4 5 6 0 1 2 3

Frequency (cycles/m) Frequency (cycles/m) Frequency (cycles/m)

Fig. 6. Time-frequency and sediment accumulation rate history results for grayscale data from the Eagle Ford Group. All analyses employ three 2p tapers and a 6-m moving window and are plotted against the adjusted depth
(bentonites and core cracks removed). A. grayscale data; B. ltered and normalized EHA results (the maximum amplitude in each 6-m window is scaled to unity, and the results are then ltered at the 80% Harmonic Condence Level;
Meyers and Hinnov, 2010). Black indicates insignicant results; C. Evolutive harmonic analysis amplitude results; D. Evolutive Power Spectral analysis results (Log10 power); E. rock accumulation rate history calculated based on tracing
the long obliquity (O1, red line). For comparison rock accumulation rate derived from bandpassed long eccentricity (black line) and ASM results (blue line) are also plotted. The ASM-derived orbital calibration is identied on gures
BeD: long eccentricity (E1), short eccentricity (E2, E3), obliquity (O1, O2) and precession (P1, P2) terms.

329
330
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344
Fig. 7. Age-depth prole of the Iona-1 core. Iona-1 core data including lithostratigraphy, lithology and Organic carbon isotope data (d13Corg); regional bentonites AeD (Elder, 1985) and X-bentonite marked. Three astronomical models
plotted along with UePb dated bentonites from the core; supplemented by regional bentonites BeD (Elder, 1985) using UePb and 40Ar/39Ar dates from Meyers et al. (2012a), acknowledging that some may represent an amalgamation of
multiple eruptions.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 331

Fig. 8. Detrended grayscale and bandpassed lter results for the ASM-derived orbital components from the grayscale of the Eagle Ford Group in Iona-1 core (plotted against the
original depth with bentonites/core crack intervals included, producing the subtle kinks and offsets in the curves): A. Detrended grayscale (light-gray curve) and 500-point moving
average (solid black); B. Precession (P1, P2); C. Obliquity (O1, O2); D. Short eccentricity (E2, E3); E. Long eccentricity (E1). Light-gray curve in b-e is the detrended 500-point moving
average curve.

The interpretation of 206Pb/238U zircon ages of bentonites within ages remain within uncertainty. However, this alternative approach
the Iona-1 core and the interpolated age model are generally should be treated with caution as it requires that the youngest
consistent within uncertainty with most of the existing 206Pb/238U dated zircon grain represents the eruption age and can be impacted
zircon and 40Ar/39Ar ages of approximately coeval bentonites from by minor post-crystallization Pb loss (e.g., zircon 6 from bentonite
the KWIS (Barker et al., 2011; Meyers et al., 2012a). Even when B7; see Fig. 5, Table 2). This situation has been recognized in a small
applying an alternative interpretation of using only the youngest number of analyses despite being greatly reduced since the
206
Pb/238U zircon ages versus the weighted mean 206Pb/238U age of development of the CA-ID-TIMs method (e.g. Davydov et al., 2010;
a statistically coherent cluster of analyses in the Iona-1 core, the Schoene et al., 2006; Meyers et al., 2012a).
332
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344
Fig. 9. Integrated regional correlation between Iona-1 and Mustang-1 cores with the USGS Portland-1 core and GSSP section, Pueblo, CO, USA. See Fig. 2 for the lithology legend; all depths are in meters, and the variable scale to reect
different sediment accumulation rates. Biostratigraphic events and isotopic events from Mustang-1 and Iona-1 (this study; Table 3 for legends) correlated with USGS Portland-1 core. Correlation also based on the short eccentricity
bandpass lter periodicity for Iona-1 (this study) and USGS Portland-1 (Meyers et al., 2001); CTB C-T Boundary. Letters A-D mark regional bentonite marker beds (after Elder, 1985). d13Corg data sources: Iona-1 and Mustang-1 (this
study), Portland-1 core (after Sageman et al., 2006; Du Vivier et al., 2014; and this study) and GSSP, Pueblo (Bowman and Bralower, 2005).
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 333

Two exceptions are observed between the 206Pb/238U zircon zircon age of the coeval bentonite in Iona-1 (bentonite D; sample
ages from Iona-1 and those from approximately coeval bentonites B3, depth 86.47 m, this study). This, therefore, supports the inter-
from the KWIS (Barker et al., 2011; Meyers et al., 2012a), namely the pretation of Meyers et al. (2012a) that all zircons from W. devonense
i) Watinoceras devonense zone bentonite (sample AZLP-08-04 from and P. exuosum zone bentonites reect complex mixtures of
Lohali Point, Arizona; see Meyers et al., 2012a) and ii) Pseudaspi- phenocrysts, xenocrysts, and detrital zircon grains and do not
doceras exuosum zone (sample AZLP-08-05 from Lohali Point, represent the youngest eruption age.
Arizona; see Meyers et al., 2012a). These bentonites were correlated For the most part, the 40Ar/39Ar ages from approximately
into the GSSP section by Kirkland (1991) corresponding to coeval bentonites from the KWIS (Meyers et al., 2012a) are slightly
Bentonite C and D of Elder (1985), respectively. The 206Pb/238U younger, but within within uncertainty of the 206Pb/238U zircon
zircon age from both bentonites when correlated into the Iona-1 age and the interpolated age model from Iona-1. Paired 40Ar/39Ar
section as bentonite C and D of Elder (1985; see Supplemental e 206Pb/238U analyses of younger Niobrara Fm bentonites from the
Information) are older beyond uncertainty than the age estimated KWIS (Sageman et al., 2014) demonstrate concordancy in the two
from interpolated astronomical age model and the 206Pb/238U systems. Feldspar 40Ar/39Ar analyses on the Iona-1 bentonites

A 93
Age (Ma)

94

95
B3 B4 B6 B7 B1B B1B*
Meyers et al.

Barker et al.

(86.47m) (100.66m) (124.78m) (127.12m) (48.40m) (48.40m)


(2012a)

(2011)

B 93
Age (Ma)

94

95
B3 B4 B6 B7 B1B B1B*
Meyers et al.

Barker et al.

(86.47m) (100.66m) (124.78m) (127.12m) (48.40m) (48.40m)


(2012a)

(2011)

Fig. 10. Interpolated age estimates for the Cenomanian-Turonian stage boundary using each of ve isotopically dated horizons in Iona-1 core as anchors: A) O1 tracing astronomical
age model; B) E1 bandpass age model. Crossed squares previous age estimates for the C-T boundary after Meyers et al. (2012a) and Barker et al. (2011); solid
squares interpolated age estimates anchored to bentonite beds in Iona-1 that are stratigraphically closest to the boundary; hollow squares interpolated age estimates from other
bentonite beds in Iona-1. Bentonites are B3; B4; B5; B6; B7 and B1B. Interpolated age estimate B1B* includes a 0.18 Ma hiatal interval at 69.6 m (see text). Error bars total
propagated uncertainties for each interpolation. Horizontal dashed line represents the interpolated age of the C-T boundary using the anchored bentonite B3 (86.47 m); dark-gray
shading analytical uncertainty of the UePb age; and light-gray total propagated uncertainty.
334 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

have yet to be performed, and should be completed to permit 5.2. Age of the Cenomanian-Turonian Boundary
paired 40Ar/39Ar e 206Pb/238U analyses. We believe the new UePb
zircon ages are sufciently precise to support the proposed age Several age estimates for the C-T stage boundary were calculated
model. through interpolation of the O1 tracing and E1 bandpass

Fig. 11. Global organic carbon isotope (d13Corg) correlation between Iona-1 (this study), KWIS reference section (Joo and Sageman, 2014), terrestrial organic matter isotope (d13CTOM)
record from Obira, Japan (Uramoto et al., 2013) and the carbonate carbon isotope (d13Ccarb) record from the English Chalk reference section (Jarvis et al., 2006). Isotope event
nomenclature follows this study, Jarvis et al. (2006), and Uramoto et al. (2013); dark-gray horizontal shading major isotopic events; light-gray horizontal shading earlier
initiation of OAE-2 CIE based on inclusion of precursor events. Isotope curves have been calibrated to the Iona-1 astronomical age model using key isotopic events. Note: two
unconformities are identied at ~92.7 Ma in Iona-1 and ~94.7 Ma in Portland-1; additional unconformities may be present in all these sections.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 335

astronomical age models using multiple UePb dated benonites model anchored using the weighted mean 206Pb/238U ages from
from the Iona-1 core as possible anchors (Fig. 10). The age estimates bentonites stratigraphically closest to the boundary (B3 at 86.47 m
include total propagated 2s uncertainties, including total isotopic and B4 at 100.66 m) are 94.10 0.13 Ma and 94.12 0.53 Ma,
uncertainty at 95% condence level (Tables 1 and 2), uncertainty in respectively (Fig. 10A). Age estimates of the C-T boundary based on
the astronomical O1 and E1 terms in quadrature (0.02 Myr and the E1 bandpass astronomical model using the same approach
0.1 Myr respectively) following methodology of Sageman et al. provide comparable estimates: 94.07 0.16 Ma and 94.21 0.54
(2014) and correlation uncertainty in the stratigraphic placement Ma, respectively (Fig. 10B). Even when interpolating weighted
of the stage boundary compared to the GSSP (2s; 0.05 Ma, see mean 206Pb/238U ages from bentonites stratigraphically furthest
supplemental information). As all the UePb ages used for the from the stage boundary (B1B: 48.4 m and B7: 127.12 m), the age
anchoring of the age model originate from the Iona-1 core, no estimates for the C-T boundary are remarkably consistent:
addition uncertainty in correlation or projection of ages is required. 93.95 0.12 Ma and 94.16 0.15 Ma (O1 tracing, Fig. 10A) and
The most condent age determination for the C-T stage boundary is 94.01 0.16 Ma and 94.20 0.18 Ma (E1 bandpass, Fig. 10B).
based on interpolation from dated bentonite beds that are strati- A small hiatus has been identied at the base of the Langtry
graphically closest to the boundary and thus have the lowest degree Member in Iona-1 (69.6 m depth; see Section 5.4) as an erosional
of additive distortion in the astronomical model and least potential surface underlying a thin gravity-ow deposit and can be correlated
for temporal gaps due to hiatuses (see Sageman et al., 2014). The C- regionally based on stable isotopes and biostratigraphy (Lowery
T boundary age estimates based on the O1 tracing astronomical age et al., 2014; Supplemental Information), and thus anchoring on the

A B

Fig. 12. BudaeEagle Ford contact interval of the Iona-1 core (152.85e153.41 m); A. white-light core photograph showing the Buda Limestone overlain by a debris ow, followed by
deposition of darker organic-rich marlstones of the Eagle Ford Group. White clasts are reworked Buda Limestone fragments delineating an unconformable; B. Ultra-violet (UV) core
photograph of same section; C. vug development within the uppermost Buda, image enlarged 3 relative to A and B.
336
A B C D E F G
Depth (m) Rock Acc. Mass Acc. TOC TOCcf Carbonate Non-Carbonate

marlstone
subsurface

limestone
Lithostrat.
Lithostrat.

bentonite
Rate Bulk Density Rate MAR MAR MAR Non-TOC MAR

Bentonite
outcrop
(g/cm3) (g/cm2/kyr) (g/cm2/kyr) (g/cm2/kyr) (g/cm2/kyr) (g/cm2/kyr)

beds
(cm/kyr)

MTD
0 1 2 3 4 5 2 2.5 3 0 5 10 15 0 0.5 1 0 50 100 0 5 10 0 5 10
0

10
Austin Chalk
Austin Chalk

20

30

40

J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344


Langtry Mb.

50
Upper Eagle Ford

60

70

80
Boquillas Fm

D
Eagle Ford Gr

90 C
100 B
A
Rock Pens Mb.

110
Lower Eagle Ford

120

130 X
KWIS Bentonites

140

150
Limestone
Limestone

160
Buda
Buda

170

180

Fig. 13. Mass accumulation rate (MAR) plots for the Iona-1 core. limestones open circles; marlstones black-lled circles; A. Lithostratigraphy/Lithology; B. Rock accumulation rates based on the O1 astronomical age model and bulk
density; density values based on wireline log and average density measurements from core plugs of representative lithologies; C. Mass accumulation rate (MAR rock accumulation rate  bulk density); D. TOC MAR; E. Carbonate-free
TOC MAR (TOCcf MAR); calculated by TOCcf MAR TOC MAR/(1-Calcite/100); F. Carbonate MAR; G. Non carbonate-non TOC MAR (MAR e TOC MAR e Carbonate MAR). Note that bentonites have been removed from calculations and
are shown in column A for reference.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 337

Plate 1. White light and SEM illustrations of selected planktic foraminifera from the Iona-1 core. All scale bars 100 mm, all images at 60x magnication. 1e2: Hedbergella
planispira, 143.37 m; 3e4: Hedbergella delrioensis, 143.37 m; 5e10: Globigerinelloides eaglefordensis, 106.94 m; 11e12: Helvetoglobotruncana praehelvetica, 79.57 m; 13: Helveto-
globotruncana praehelvetica-helvetica transition, 79.57 m; 14: Helvetoglobotruncana praehelvetica, 74.25 m; 15e16: Whiteinella archaeocretacea, 130.74 m; 17e18: Rotalipora mon-
tsalvensis, 109.22 m.
338 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Plate 2. White light and SEM illustrations of selected planktic foraminifera from the Iona-1 core. All scale bars 100 mm, all images at 60x magnication. 1: Favusella washitensis,
171.56 m; 2e3: Heterohelix globulosa, 98.08 m; 4e5: Heterohelix moremani, 98.08 m; 6e7: Anaticinella multiloculata, 109.22 m; 8: Rotalipora cf. deekei, 132.76 m; 9e10: Thalmaninella
appenninica, 143.37 m; 11e14: Rotalipora cushmani, 107.82 m; 15e16: Dicarinella hagni, 104.00 m; 17e18: Thalmaninella greenhornensis, 107.82 m.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 339

shallowest bentonite B1B at 48.4 m depth, probably results in an the duration of the OAE-2 interval would be substantially longer
underestimate, with the age of the C-T boundary being older due to a (0.92 0.17 Myrs) than other KWIS records and that previously
temporal gap in the astronomical model. Following the approach of reported from European sections (~0.45 Myrs; Voigt et al., 2008)
Sageman et al. (2014) the weighted mean 206Pb/238U bentonite ages and suggest an earlier initiation of the OAE-2 than previously
below the unconformity and the astronomical age model were used recognized. It is more likely that the LO of T. greenhornensis is
to interpolate the 48.4 m bentonite age. This approach results in an stratigraphically diachronous between Europe and the KWIS, as is
offset of approximately 0.15e0.21 Myr between the UePb age and the case with other marker species, such as the ammonite Wati-
the interpolated age, which may represent the duration of the hiatus noceras devonense (see discussion in Voigt et al., 2008). Therefore,
at 69.6 m depth. When the estimated duration of the unconformity our preferred interpretation is that the three precursor isotope
is added to the age estimate for the C-T boundary using the events identied in Iona-1, USGS Portland-1 and the GSSP, Pueblo
bentonite anchor at 48.4 m, the resulting age using the O1 tracing (Fig. 9) occur below the onset of the OAE-2 interval as previously
and E1 bandpass astronomical models are 94.10 0.33 Ma and recognized and are forerunners to the main CIE. The occurrence of
94.19 0.35 Ma, respectively (including propagation of total errors) these precursor events in numerous locations within the KWIS
and are both consistent with the ages interpolated using other an- (e.g., Iona-1; USGS Portland-1; GSSP, Pueblo) also correspond with
chors (Fig. 10). other global records where they occur below the rst rise in the
Our age estimates for the C-T boundary are, therefore, consistent stable carbon isotope values, but where their signicance has not
with that of 94.12 0.13 previously proposed by Barker et al. (2011) been highlighted; e.g., the New Jersey Shelf (van Helmond et al.,
and within uncertainty of the 93.9 0.15 Ma proposed by Meyers 2014), Equatorial Atlantic (Forster et al., 2007), South Atlantic
et al. (2012a; Fig. 10). (Forster et al., 2008), and European sections such as the Vocantian
Basin, France (Jarvis et al., 2011) and Wunstorf, Germany (Du Vivier
5.3. Age and duration Cenomanian and Turonian Isotopic Events et al., 2014). It is possible that these precusor events are part of the
OE-2 CIE interval and requires further studies on complete strati-
Previous astronomical age calibrations of the Middle Cen- graphic sections.
omanian Event (MCE) are limited to those of Lanci et al. (2010) and Three Turonian isotope events are also expressed in our d13Corg
Mitchell et al. (2008) who conducted analyses on the Scaglia record (Fig. 2, Table 3) that closely resemble positive carbon isotope
Bianca Fm, at Furlo, Italy. In the KWIS, Ma et al. (2014) provided an events from terrestrial organic matter (d13CTOM) within the Turo-
astronomical age model down to the Lincoln Limestone Member nian Yezo Group, Japan (YT1-YT3; Uramoto et al., 2013). These
of the Greenhorn Fm., but did not extend their analysis down to organic carbon events can be correlated with reference carbon
the Graneros Shale where the MCE has been identied (Joo and isotope trends based on bulk carbonate (d13Ccarb) from Europe
Sageman, 2014). Therefore, the astronomical age model pre- (Jarvis et al., 2006; Wendler, 2013; see Uramoto et al., 2013) and the
sented from Iona-1 provides an important calibration point for the Western Interior Seaway (Sageman et al., 2006; Joo and Sageman,
MCE and is expressed in the d13Corg record occurring within the 2014). In particular, the early to middle Turonian CIE in Iona-1
highest TOC interval near the base of the Eagle Ford Gr corresponds with the YT1 Yezo Group and Lulworth-Round Down
(146e148 m). Based on the O1 astronomical age model, the MCE excursions in the English Chalk reference section (Jarvis et al.,
interval in Iona-1 spans 96.57 0.12 Ma to 96.36 0.12 Ma (see 2006), which also encompass the higher frequency Tu7eTu10
Table 3 for details). isotope events (Voigt et al., 2007; Wendler, 2013; Fig. 11). Based on
Estimates of the duration of the C-T CIE which denes OAE-2 the tracing of the orbital O1 signal age model, the early-middle
depend on the denition of its onset, termination, and complete- Turonian CIE is 93.13 0.12 to 92.49 0.12 Ma, with a duration
ness of the sedimentary record. The termination is particularly of ~0.64 0.17 Ma.
difcult to constrain as the return of isotope values to background The middle Turonian event identied in Iona-1 corresponds to
levels is gradual. This is further complicated as carbon isotope re- the d13CTOM YT-2 event in Japan (Uramoto et al., 2013) as well as
cords are impacted by factors such as: i) the relative abundance of with the Pewsey d13Ccarb Event in the English Chalk (Jarvis et al.,
mixed components with different d13C isotopic values; ii) isotope 2006) and Western Interior Seaway (Joo and Sageman, 2014). It is
fractionation due to physiological change and depositional envi- calibrated in the Iona-1 core to initiate at 91.95 0.12 Ma (O1
ronments; and iii) d13C isotopic values of the carbon source, leading tracing), with a duration of ~0.16 Myr. The shift towards negative
to offsets from the global exogenic signatures (Sluijs and Dickens, d13Corg values in the upper Eagle Ford and lowermost Austin
2012; Wendler, 2013). However, even though these factors affect Chalk in Iona-1 is correlated to the d13CTOM YT-3 event in Japan
the shape and magnitude of the d13C isotopic curve, they should not (Uramoto et al., 2013) and, based on our astronomically tuned
impact the timing or duration of these globally recognized events age model, corresponds with the latest Turonian to earliest
(see Wendler, 2013). Using the O1 orbital tracing astronomical age Coniacian with the initial increase in d13Corg values at
model, the duration of the OAE-2 interval in Iona-1 is calculated 90.83 0.12 Ma. The late Turonian-Coniacian d13Corg event
based on the base prominent excursion (106.94 m) to the end of the identied in Iona-1 shows good correspondence with the
plateau in d13Corg values (92.73 m) as 0.71 0.17 Myr. A comparison d13CTOM YT-3 event in Japan (Uramoto et al., 2013) and the En-
of the bandpass-ltered short eccentricity cycle from Iona-1 and glish Chalk reference section (Jarvis et al., 2006; Voigt et al.,
the USGS Portland core (Meyers et al., 2012a; Sageman et al., 2006; 2007; Wendler, 2013). Specically, the positive excursion may
Meyers et al., 2001); shows remarkable similarity between the two correspond with the Hitchwood event, whereas negative ex-
records (Fig. 9). The duration of OAE-2 is consistent with previous cursions may correspond with the Bridgewick and Navigation
estimates from the KWIS, particularly the USGS Portland core events. In conclusion, the correlation between available marine
(0.56e0.67 Myr; Sageman et al., 2006) and Aristocrat Angus core d13Corg record in the KWIS of North America, with d13CTOM re-
(0.52e0.62 Myr; Ma et al., 2014) despite the presence of an un- cords in Japan (Uramoto et al., 2013) and the European d13Ccarb
conformity at the base of the Bridge Creek Limestone Member in reference prole (Jarvis et al., 2006; Wendler, 2013; Voigt et al.,
the USGS Portland core. However, if onset of OAE-2 was taken at the 2007) supports a linkage between terrestrial and marine car-
rst possible rise of the stable carbon isotope curve (112.45 m) bon reservoirs within Late Cretaceous global ocean-atmosphere-
including the precursor events and occurring below the LO of biosphere systems, as suggested by Barclay et al. (2010), Uramoto
T. greenhornensis as also reported in Europe (Jarvis et al., 2011), then et al. (2013) and Joo and Sageman (2014).
340 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Plate 3. Selected nannofossil marker taxa from the Iona-1 core. All images to same scale bar (top left image). 1. Stoverius achylosus, 18.70e18.67 m, XPL; 2. Stoverius achylosus,
18.70e18.67 m, PPL; 3. Marthasterites furcatus, 18.70e18.67 m, XPL; 4. Marthasterites furcatus, 18.70e18.67 m, XPL, Same specimen. 45 ; 5. Marthasterites furcatus (fragment),
18.70e18.67 m, XPL; 6. Marthasterites furcatus, 18.70e18.67 m, XPL, Same specimen. 45 ; 7. Eiffellithus eximius, 18.70e18.67 m, XPL; 8. Eiffellithus eximius, 18.70e18.67 m, XPL, Same
specimen. 45 ; 9. Eiffellithus turriseiffelii, 127.77 m, XPL; 10. Eiffellithus turriseiffelii, 127.77 m, XPL, Same specimen. 45 ; 11. Eprolithus oralis, 81.40 m, XPL; 12. Eprolithus eptapetalus,
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 341

5.4. Accumulation rates and lithology inorganic and organic matter burial rates and our understanding of
local, regional, and global perturbations to the carbon cycle and
The rock accumulation rate for the lower Cenomanian Buda KWIS depositional systems during the Late Cretaceous. In partic-
Limestone is calculated by linear extrapolation between two ular, our record from the Maverick Sub-basin from a starved, distal
available bentonite UePb zircon ages and is relatively high part of the KWIS demonstrates that the maximum organic carbon
(~4e6 cm/ka). The contact between the Buda Limestone and Eagle accumulation and preservation occurred during the early to middle
Ford Group is erosional and unconformable. Based on the Cenomanian (95e97 Ma), which does not correspond with the
biostratigraphic data, the hiatus between the two formations is global peak in organic-matter sequestration at the C-T boundary
constrained to the early Cenomanian, and a linear extrapolation indicated by the positive global CIE. Similarly, the recorded
between geochronological events and the astronomical age model decrease in carbon preservation during the OAE-2 interval in Iona-1
indicate a duration of ~0.41 Myr with resumption of deposition (Eldrett et al., 2014) is also observed further to the north in the
marking the base of the Eagle Ford Group occurring at KWIS (e.g. Sageman and Lyons, 2003). This indicates that massive
97.05 0.12 Ma. This event might correspond with a eustatic sea- carbon sequestration occurred in some other region than the KWIS
level fall (Ce2- Ce3 sequence boundary of Haq et al., 1987) as nez
(e.g., Kuypers et al., 2004; Trabucho Alexandre et al., 2010; Jime
expressed by paleokarst in the middle East and North Africa (e.g., Berrocoso et al., 2010; Eldrett et al., 2014). Carbonate mass accu-
Wilmsen et al., 2013). In Iona-1, the uppermost few meters of the mulation rates are highest in both marl and limestone lithologies in
Buda Limestone contains vugs with calcite spar that also may the underlying Buda Limestone and overlying Austin Chalk sug-
represent exposure-related karst during the early Cenomanian gesting relatively reduced carbonate productivity during Eagle Ford
implying subaerial exposure before the deposition of the marine deposition. This reects the unique environmental and biological
sediment of Eagle Ford Group. However, they may alternatively conditions associated with the organic-rich Eagle Ford in contrast
result from more recent meteoric water interactions, once both to the organic-poor bounding units. A more detailed discussion of
formations were exhumed (Fig. 12). Tilted blocks of well-bedded the astronomical and climate relationships with lithofacies are
Eagle Ford Group collapsed within the karstic openings of the contained in a companion paper (Eldrett et al., 2015).
Buda Limestone are also observed ~90 km northwest of the Iona-1
core location where the Eagle Ford Group crops out at the surface 6. Conclusions
along US Highway 90 (pers. obs., 2012) thus favoring the hypothesis
that the top of the Buda Limestone is the result of meteoric karst We have developed two relatively continuous astronomically
rather than exposure-related karst, although further work is tuned age models for the Cenomanian, Turonian and earliest Con-
required to constrain the timing of meteoric karstication. iacian from a research core located in the distal and clastic sedi-
Rock accumulation rates obtained from O1 astronomical age ment starved part of the southern gateway of the Cretaceous
model for the section above the gravity-ow deposits in the Western Interior Seaway. These analyses provide a reliable tem-
lowermost 2e6 m of the Eagle Ford Group indicate fairly constant poral framework for interpreting the dynamics of environmental
deposition throughout the 7 Myr represented by the studied in- change throughout a period of major perturbations to the global
terval (Table 4). The lower Eagle Ford has rock accumulation rates carbon cycle. This core is one of the most complete and best-
between 1.5 and 1.9 cm/ka, which slightly increase at the lower- preserved global records of the Cenomanian, Turonian and early
upper Eagle Ford transition (2.4 cm/ka), then decrease in the up- Coniacian stages, and provides insights into the nature of the
per Eagle Ford (1.0e2.0 cm/ka) with the lowest rates in the Langtry associated Greenhouse conditions in a mid-latitude, shelf-platform
Member near the top of the Upper Eagle Ford. Rock accumulation setting.
rates increase dramatically in the overlying Austin Chalk (3.8 cm/ By capitalizing on the abundance of the most important marine
ka). Individual bentonite layers would have been deposited at microfossil groups e calcareous nannofossils; planktic and benthic
accumulation rates >100e1000 cm/ka and were, therefore, foraminifers, radiolaria and dinoagellate cysts e we have identi-
excluded from Fig. 13. ed numerous age diagnostic events which have been calibrated to
Middle to upper Turonian strata in the KWIS are generally an astronomical age model, incorporating UePb ID-TIMS zircon
missing due to an inferred sequence boundary representing a ages from bentonites, together providing high-precision indepen-
regional regressive event (see Roberts and Kirschbaum, 1995). This dent age control for regional and global correlations. We have also
stratigraphic interval analyzed in the Iona-1 core shows that within identied ve major marine isotopic events occurring in the middle
the resolution of our astronomical age model, the record is strati- Cenomanian (MCE); C-T boundary interval (OAE-2); early-middle
graphically continuous, with the exception of one minor middle Turonian (EMTE); middle Turonian (MTE) and late Turonian-
Turonian interval at 92.71 0.12 Ma (69.6 m) that shows a short Coniacian (LTCE); these events have been calibrated to the astro-
hiatus with a duration of ~0.18 0.03 Myr as expressed in the core nomical age models and correlate with similar isotopic events in
as an erosional surface underlying a thin gravity-ow deposit. This the terrestrial record. We have further constrained the duration of
submarine hiatus preserved in Iona-1 may represent the distal OAE-2 to ~0.71 Myrs, consistent with other records from the KWIS,
product of the shallower-water slumping events observed in the and have identied several precursor events leading up to the main
Trans-Pecos region (pers. obs., 2012). CIE which globally correlate. Furthermore, our integrated astro-
The astronomically tuned age model proposed for the Cen- nomical age models indicate a C-T boundary age of 94.10 0.13 Ma
omanian, Turonian and earliest Coniacian stages enables the or 94.07 0.16 Ma consistent with that proposed by Barker et al.
reconstruction of mass accumulation rates for the main sedimen- (2011) for the Western Canada Foreland Basin and within uncer-
tary components (Fig. 13), which have important implications for tainty of that proposed for the GSSP by Meyers et al. (2012a).

81.40 m, XPL; 13. Eprolithus eptapetalus, 81.40 m, XPL; 14. Radiolithus orbiculatus, 81.40 m, XPL; 15. Radiolithus planus, 81.40 m, XPL; 16. Quadrum gartneri, 41.89 m, XPL; 17. Quadrum
gartneri, 41.89 m, XPL, Same specimen. 45 ; 18. Quadrum intermedium, 41.89 m, XPL; 19. Quadrum intermedium, 41.89 m, XPL, Same specimen. 45 ; 20. Quadrum eneabrachium,
25.93 m, XPL; 21. Rhagodiscus asper, 81.40 m, XPL; 22. Rhagodiscus asper, 81.40 m, XPL, Same specimen. 45 ; 23. Axopodorhabdus albianus, 119.49 m, XPL; 24. Axopodorhabdus
albianus, 119.49 m, XPL, Same specimen. 45 ; 25. Helenea chiastia, 119.49 m, XPL; 26. Helenea chiastia, 119.49 m, XPL, Same specimen. 45 ; 27. Cretarhabdus striatus, 121.09 m, XPL;
28. Cretarhabdus striatus, 121.09 m, XPL, Same specimen. 45 ; 29. Corollithion kennedyi, 112.77 m, XPL; 30. Corollithion kennedyi, 112.77 m, XPL, Same specimen. 45 .
342 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Plate 4. Selected dinocyst taxa from Iona-1; All gures are to the same scale (see 50 mm scale bar). 1. Cyclonephelium membraniphorum, Iona-1, 86.65 m, V30/2; 2. Senoniasphaera
rotundata, Iona-1, 7.91 m, T27/4; 3. Senoniasphaera turonica, Iona-1, 7.91 m, J28/0; 4. Heterosphaeridium difcile, Iona-1, 45.72 m, X18/2; 5. Isabelidinium magnum, Iona-1, 7.91 m,
W36/4; 6. Senoniasphaera turonica, Iona-1, 52.57 m, U21/1; 7. Eurydinium glomeratum, Iona-1, 83.22 m, X25/2; 8. Litosphaeridium siphoniphorum, Iona-1, 119.77 m, F32/0; 9.
Cyclonephelium longispinatum, Iona-1, 119.77 m, P23/1; 10. Bosedinia spp. cf. B. sp. 1 & sp. 3 of Prauss (2012b), Iona-1, 127.31 m, V25/2; 11. Bosedinia spp. cf. B. sp.1 & sp.3 of Prauss
(2012b), Iona-1 T/S, 260.81 m, V23/0; 12. Adnatosphaeridium tutulosum, Iona-1, 104.18 m, N17/0; 13. Adnatosphaeridium? chonetum, Iona-1, 88.77 m, V20/2.
J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344 343

Acknowledgments Cycles: Orbital forcing of organic-rich sedimentary rocks from the Western
Interior Seaway, USA. Earth Planetary Science Letters 423, 98e113. http://
dx.doi.org/10.1016/j.epsl.2015.04.026.
This research used samples from USGS Portland-1 core provided Fisher, C.G., Arthur, M.A., 2002. Water mass characteristics in the Cenomanian US
by the Core Research Center, Denver. We acknowledge analytical Western Interior Seaway as indicated by stable isotopes of calcareous organ-
contributions by Bob Gregory (the Stable Isotope Laboratory at isms. Palaeogeoghraphy, Palaeoclimatology, Palaeoecology 188, 189e213.
Forster, A., Schouten, S., Moriya, K., Wilson, P.A., Sinninghe Damste , J.S., 2007. Tropical
Southern Methodist University, USA) and colleagues at Shell, in warming and intermittent cooling during the Cenomanian/Turonian oceanic
particular other members of the mudrock team and ASAP team anoxic event 2: sea surface temperature records from the equatorial Atlantic.
including Andy Bishop, Katrin Ruckwied, Sarah-Jane Jackett, Rob Paleoceanography 22, PA1219. http://dx.doi.org/10.1029/2006PA001349.
Forster, A., Kuypers, M.M.M., Turgeon, S.C., Brumsack, H.-J., Petrizzo, M.R., Sinninghe
Campbell, Rui da Gama, and Iain Prince for internal review. We also Damste , J.S., 2008. The Cenomanian/Turonian oceanic anoxic event in the South
thank Aaron Shunk and Kyle Patterson for astrochronology dis- Atlantic: new insights from a geochemical study of DSDP Site 530A. Palae-
cussions as well as David Lavallee and Nick Howes for writing ogeography, Palaeoclimatology, Palaeoecology 267, 256e283.
Haq, B.U., Hardenbol, J., Vail, P.R., 1987. Chronology of uctuating sea level since the
MatLab code enabling the isolation and removal of core cracks from Triassic. Science 235, 1156e1167.
the grayscale data. We acknowledge Malcolm Jones and Palyno- Hinnov, L.A., Hilgen, F.J., 2012. Cyclostratigraphy and Astrochronology. In:
logical Laboratory Services (PLS) Ltd for excellent palynological Gradstein, F.M., Ogg, J.G., Schmitz, M.D., et al. (Eds.), The Geologic Time Scale 2012.
Elsevier, Boston, USA. http://dx.doi.org/10.1016/B978-0-444-59425-9.00004-4.
preparations. We thank Brad Sageman, Northwestern Univ., USA for Hoke, G.D., Schmitz, M.D., Bowring, S.A., 2014. An ultrasonic method for isolating
providing USGS Portland-1 core d13Corg data for the Bridge Creek nonclay components from clay-rich material. Geochemistry Geophysics Geo-
Limestone Member and Ian Jarvis, Kingston Univ., UK for stable systems 15, 492e498. http://dx.doi.org/10.1002/2013GC005125.
Immenhauser, A., Holmden, C., Patterson, W.P., 2008. Interpreting the carbon
isotope data from the UK reference section. We would also like to
isotope record of ancient epeiric seas: lessons from the recent. In: Pratt, B.R.,
thank Stephen Meyers and Andrew Gale for external peer reviews Holmden, C. (Eds.), Dynamics of Epeiric Seas. Geological Association of Canada,
that greatly improved the manuscript. This research was conducted Special Paper, pp. 135e174.
and funded by Shell International Exporation and Production Inc., Irwin, H., Curtis, C., Coleman, M., 1977. Isotopic evidence for source of diagenetic
carbonates formed during burial of organic-rich sediments. Nature 269, 209e213.
and we thank our Shell leadership for permission to publish. Jaffey, A.H., Flynn, K.F., Glendenin, L.E., Bently, W.C., Essling, A.M., 1971. Precision
measurements of half-lives and specic activities of 235U and 238U. Physical
Review C: Nuclear Physics 4, 1889e1906.
References Jarvis, I., Gale, A.S., Jenkyns, H.C., Pearce, M.A., 2006. Secular variation in Late
Cretaceous carbon isotopes: a new d13C reference curve for the Cenomanian-
Barker, I.R., Moser, D.E., Kamo, S.L., Plint, G.A., 2011. High-precision UePb zircon Santonian (99.6e83.5 Ma). Geological Magazine 143, 561e608.
IDeTIMS dating of two regionally extensive bentonites: Cenomanian Stage, Jarvis, I., Lignum, J.S., Gro cke, D.R., Jenkyns, H.C., Pearce, M., 2011. Black shale
Western Canada Foreland Basin. Canadian Journal Earth Science 48, 543e556. deposition, atmospheric CO2 drawdown, and cooling during the Cenomanian-
http://dx.doi.org/10.1139/E10-042. Turonian Oceanic Anoxic Event. Paleoceanography 26, PA3201. http://
Barclay, R.S., McElwain, J.C., Sageman, B.B., 2010. Carbon sequestration activated by dx.doi.org/10.1029/2010PA002081.
a volcanic CO2 pulse during Ocean Anoxic Event 2. Nature Geosciences 3, Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochemistry
205e208. http://dx.doi.org/10.1038/ngeo757. Geophysics Geosystems v. 11, Q03004. http://dx.doi.org/10.1029/2009GC002788.
Bowman, A.R., Bralower, T.J., 2005. Paleoceanographic signicance of high resolution Jimenez Berrocoso, A., MacLeod, K.G., Martin, E.E., Bourbon, E., London ~ o, C.I., Basak, C.,
carbon isotope records across the Cenomanian-Turonian boundary in the Western 2010. Nutrient trap for Late Cretaceous organic-rich black shales in the tropical
Interior and New Jersey coastal plain, USA. Marine Geology 217, 305e321. North Atlantic. Geology 38, 1111e1114. http://dx.doi.org/10.1130/G31195.1.
Bowring, J.F., McLean, N.M., Bowring, S.A., 2011. Engineering cyber infrastructure for Joo, Y.J., Sageman, B.B., 2014. Cenomanian to Campanian carbon isotope chemo-
U-Pb geochronology: Tripoli and U-Pb_Redux. Geochemistry, Geophysics, stratigraphy from the Western Interior Basin, U.S.A. Journal of Sedimentary
Geosystems 12, Q0AA19. http://dx.doi.org/10.1029/2010GC003479. Research 84, 529e542. http://dx.doi.org/10.2110/jsr.2014.38.
Burnett, J.A., 1998. Upper Cretaceous. In: Bown, P.R. (Ed.), Calcareous Nannofossil Kennedy, W.J., Walaszczyk, I., Cobban, W.A., 2000. Pueblo, Colorado, USA, candidate
Biostratigraphy. Chapman and Hall, London, pp. 132e199. global boundary stratotype section and point for the base of the Turonian stage
Caron, M., Robaszynski, F., Amedro, F., Baudin, F., Deconinck, J., Hochuli, P.A., Nielsen, K.S., of the Cretaceous, and for the base of the Middle Turonian substage, with a
Tribovillard, N., 1999. Estimation de la duree de l'evenement anoxique global au revision of the Inoceramidae (Bivalvia). Acta Geologica Polonica 50, 295e334.
passage Ce nomanien/Turonien: approche cyclostratigraphique dans la formation Kenndey, W.J., Walaszczyk, I., Cobban, W.A., Dodsworth, P., Elder, W.P., Gale, A.S.,
Bahloul en Tunisie central. Bulletin de la Societe Geologique de France 170, 145e160. Scott, G.R., Hancock, J.M., Voigt, S., Kirkland, J.I., 2005. The Global Boundary
Cherniak, D.J., Watson, E.B., 2003. Diffusion in Zircon. In: Reviews in Mineralogy, v. Stratotype Section and Point for the base of the Turonian Stage of the Creta-
53, pp. 113e145 ch. 5. ceous. Colorado, Pueblo, pp. 93e104. U.S.A. Episodes, 28.
Condon, D., Schoene, B., Bowring, S.A., Parrish, R., McLean, N., Noble, S., Crowley, Q., Kirkland, J.1, 1991. Lithostratigraphic and biostratigraphic framework for the Mancos
2007 [abs.]. EARTHTIME: Isotopic Tracers and Optimized Solutions for High- Shale (late Cenomanian to middle Turonian) at Black Mesa, northeastern Arizona.
precision U-Pb ID-TIMS Geochronology, v. 88. Eos (Transactions, American In: Nations, J.D., Eaton, J.G. (Eds.), Stratigraphy, Depositional Environments, and
Geophysical Union). no. 52, V41E-06. Sedimentary Tectonics of the Southwestern Margin Cretaceous Western Interior
Davydov, V.I., Crowley, J.L., Schmitz, M.D., Poletaev, V.I., 2010. High-precision Seaway, Geological Society of America Special Paper 260, pp. 85e111.
UePbzircon age calibration of the global Carboniferous time scale and Milan- Krogh, T.E., 1973. A low contamination method for hydrothermal decomposition of
kovitch-band cyclicity in the Donets Basin, eastern Ukraine. Geochemistry zircon and extraction of U and Pb for isotopic age determinations. Geochimica
Geophysics Geosystems 11 (1). http://dx.doi.org/10.1029/2009GC002736. et Cosmochimica Acta 37, 485e494.
Dodsworth, P., 1995. A note of caution concerning the application of quantitative Kuhnt, W., Nederbragt, A., Leine, L., 1997. Cyclicity of Cenomanian-Turonian organic-
palynological data from oxidized preparations. Journal of Micropaleontology 14, 6. carbon-rich sediments in the Tarfaya Atlantic Coastal Basin (Morocco). Creta-
Dodsworth, P., 1996. Stratigraphy, microfossils and depositional environments of ceous Research 18, 587e601.
the lowermost part of the Welton Chalk Formation (late Cenomanian to early Kuhnt, W., Luderer, F., Nederbragt, S., Thurow, J., Wagner, T., 2004. Orbital-scale
Turonian, Cretaceous) in eastern England. Proceedings of the Yorkshire record of the late Cenomanian-Turonian oceanic anoxic event (OAE-2) in the
Geological Society 51, 45e64. Tarfaya Basin (Morocco). International Journal Earth Science 94, 147e159.
Dodsworth, P., 2000. Trans-Atlantic dinoagellate cyst stratigraphy across the Kuiper, K.F., Deino, A., Hilgen, F.J., Krijgsman, W., Renne, P.R., Wijbrans, J.R., 2008.
Cenomanian-Turonian (Cretaceous) Stage boundary. Journal of Micropaleon- Synchronizing rock clocks of Earth history. Science 320, 500e504. http://
tology 19, 69e84. dx.doi.org/10.1126/science, 1154339.
Dodsworth, P., 2004. The palynology of the CenomanianeTuronian (Cretaceous) Kuypers, M.M.M., Lourens, L.J., Rijpstra, W.I.C., Pancost, R.D., Nijenhuis, I.A., Sinninghe
boundary succession at Aksudere in Crimea, Ukraine. Palynology 28, 129e141. Damste, J.S., 2004. Orbital forcing of organic carbon burial in the proto North Atlantic
Du Vivier, A.D.C., Selby, D., Sageman, B.B., Jarvis, I., Gro cke, D.R., Voigt, S., 2014. during oceanic anoxic event 2. Earth and Planetary Science Letters 228, 465e482.
Marine 187Os/188Os isotope stratigraphy reveals the interaction of volcanism Lanci, L., Muttoni, G., Erba, E., 2010. Astronomical tuning of the Cenomanian Scaglia
and ocean circulation during Oceanic Anoxic Event 2. Earth Planetary Science Bianca Formation at Furlo, Italy. Earth and Planetary Science Letters 292,
Letters v. 389, 23e32. 231e237. http://dx.doi.org/10.1016/j.epsl.2010.01.041.
Elder, W.P., 1985. Biotic patterns across the Cenomanian-Turonian extinction Laskar, J., Robutel, P., Joutel, F., Gastineau, M., Correia, A.C.M., Levrard, B., 2004.
boundary near Pueblo, Colorado. In: Pratt, L.M., Kauffman, E.G., Zelt, F.B. (Eds.), A long-term numerical solution for the insolation quantities of the Earth,. As-
Field Trip Guidebook No. 4, 157e169. tronomy and Astrophysics 428, 261e285.
Eldrett, J.S., Minisini, D., Bergman, S.C., 2014. Decoupling of the carbon cycle during Laskar, J., Fienga, A., Gastineau, M., Manche, H., 2011. La2010: A new orbital solution
Ocean Anoxic Event-2. Geology v. 42, 567e570. http://dx.doi.org/10.1130/ for the long-term motion of the Earth. Astronomy & Astrophysics 532, A89.
G35520.1 no. 7. Laurin, J., Meyers, S.R., Ulicny, D., Jarvis, I., Sageman, B., 2015. Axial-obliquity control
Eldrett, J.S., Ma, C., Ozkan, A., Bergman, S.C., Minisini, D., Lutz, B., Macaulay, C., on the greenhouse carbon budget through middle- to high-latitude reservoirs.
Jackett, S.-J., Kelly, A.E., 2015. Origin of Upper Cretaceous Limestone-Marl Paleoceanography. http://dx.doi.org/10.1002/2014PA002736.
344 J.S. Eldrett et al. / Cretaceous Research 56 (2015) 316e344

Locklair, R.E., Sageman, B.B., 2008. Cyclostratigraphy of the Upper Cretaceous Nio- Renne, P.R., 2014. Some footnotes to the optimization-based calibration of the
40
brara Formation, Western Interior, U.S.A.: A Coniacian-Santonian orbital time- Ar/39Ar system. Geological Society, London, Special Publications. http://
scale. Earth and Planetary Science Letters 269, 539e552. dx.doi.org/10.1144/SP378.17, 2014, 378:21e31, rst published on July 25, 2013.
Lowery, C.M., Corbett, M.J., Leckie, M., Watkins, D., Romero, A.M., Pramudito, A., Renne, P.R., Mundil, R., Balco, G., Min, K., Ludwig, K.R., 2010. Joint determination of
40
2014. Foraminiferal and nannofossil paleoecology and paleoceanography of the K decay constants and 40Ar*/40K for the Fish Canyon sanidine standard, and
CenomanianeTuronian Eagle Ford Shale of southern Texas. Palaeogeography, improved accuracy for 40Ar/39Ar geochronology. Geochimica et Cosmochimica
Palaeoclimatology, Palaeoecology 413, 49e65. http://dx.doi.org/10.1016/ Acta v. 74, 5349e5367. http://dx.doi.org/10.1016/j.gca.2010.06.017.
j.palaeo.2014.07.025. Ricken, W., 1986. Diagenetic Bedding: a Model for LimestoneeMarl Alternations.
Ludwig, K.R., 2003. User's manual for Isoplot/Excel, version 3.0: A geochronological Springer Verlag, Berlin, p. 210.
toolkit for Microsoft Excel, Special Publication 4. Berkeley Geochronology Roberts, L.N.R., Kirschbaum, M.A., 1995. Paleogeography of the Late Cretaceous of
Center, Berkeley, California. the Western Interior of middle North America e Coal distribution and sediment
Ma, C., Meyers, S.R., Sageman, B.B., Singer, B.S., Jicha, B.R., 2014. Testing the astro- accumulation. U.S. Geological Survey Professional Paper 1561, p. 115.
nomical time scale for Oceanic Anoxic Event 2, and its extension into Cen- Sageman, B.B., Lyons, T.W., 2003. Geochemistry of ne-grained sediments and
omanian Strata of the Western Interior Basin (U.S.A.). Geological Society of sedimentary rocks. In: Mackenzie, F. (Ed.), Treatise on Geochemistry, v. 7.
America Bulletin 126, 974e989. Elsevier, New York, pp. 115e158.
MacLeod, K.G., Huber, B.T., Berrocoso, A.J., Wendler, I., 2013. A stable and hot Turonian Sageman, B.B., Singer, B.S., Meyers, S.R., Siewert, S.E., Walaszczyk, I., Condon, D.J.,
without glacial d18O excursions is indicated by exquisitely preserved Tanzanian Jicha, B.R., Obradovitch, J.D., Sawyer, D.A., 2014. Integrating 40Ar/39Ar, U-Pb, and
foraminifera. Geology 41, 1083e1086. http://dx.doi.org/10.1130/G34510.1. astronomical clocks in the Cretaceous Niobrara Formation, Western Interior
Malinverno, A., Erba, E., Herbert, T.D., 2010. Orbital tuning as an inverse problem: Basin, USA. GSA Bulletin 126, 956e973. http://dx.doi.org/10.1130/B30929.1.
Chronology of the early Aptian oceanic anoxic event 1a (Selli level) in the Sageman, B.B., Meyers, S.R., Arthur, M.A., 2006. Orbital timescale for the
Cismon APTICORE. Paleoceanography 25, PA2203. Cenomanian-Turonian boundary stratotype and OAE II, central Colorado, USA.
Mattinson, J.M., 2005. Zircon UePb chemical abrasion (CA-TIMS) method: Com- Geology 34, 125e128.
bined annealing and multi-step partial dissolution analysis for improved pre- Sageman, B.B., Rich, J., Arthur, M.A., Bircheld, G.E., Dean, W.E., 1997. Evidence for
cision and accuracy of zircon ages. Chemical Geology 220, 47e66. http:// Milankovitch periodicities in Cenomanian-Turonian lithologic and geochemical
dx.doi.org/10.1016/j.chemgeo.2005.03.011. cycles, Western Interior U. S. Journal of Sedimentary Research 67, 285e301.
Mattinson, J.M., 2010. Analysis of the relative decay constants of 235U and 238U by Schlanger, S.O., Jenkyns, H.C., 1976. Cretaceous Oceanic Anoxic Events: Causes and
multi-step CA-TIMS measurements of closed-system natural zircon samples. Consequences. Geologie en Mijnbouw 55, 179e184.
Chemical Geology 275, 186e198. http://dx.doi.org/10.1016/j.chemgeo.2010.05.007. Schrag, D.P., Higgins, J.A., Macdonald, F.A., Johnston, D.T., 2013. Authigenic car-
Melim, L.A., Westphal, H., Swart, P.K., Eberli, G.P., Munnecke, A., 2002. Questioning bonate and the history of the global carbon cycle. Science 339 (6119), 540e543.
carbonate diagenetic paradigms: evidence from the Neogene of the Bahamas. Schmitz, M.D., Bowring, S.A., 2001. UePb zircon and titanite systematic of the Fish
Marine Geology 185, 27e53. Canyon Tuff: an assessment of high-precision UePb geochronology and its
Mertens, K.N., Verhoeven, K., Verleye, T., Louwye, S., Amorim, A., Ribeiro, S., application to young volcanic rocks. Geochimica et Cosmochimica Acta 65 (15),
Deaf, A.S., Harding, I.C., De Schepper, S., Gonz alez, C., Kodrans-Nsiah, M., De 2571e2587. http://dx.doi.org/10.1016/S0016-7037(01)00616-0.
Vernal, A., Henry, M., Radi, T., Dybkjaer, K., Poulsen, N.E., Feist-Burkhardt, S., Schoene, B., Crowley, J., Condon, D., Schmitz, M., Bowring, S., 2006. Reassessing the
Chitolie, J., Heilmann-Clausen, C., Londeix, L., Turon, J.-L., Marret, F., uranium decay constants for geochronology using IDTIMS U-Pb data. Geo-
Matthiessen, J., McCarthy, F.M.G., Prasad, V., Pospelova, V., Kyfn Hughes, J.E., chimica et Cosmochimica Acta 70 (2), 426e445.
Riding, J.B., Rochon, A., Sangiorgi, F., Welters, N., Sinclair, N., Thun, C., Sluijs, A., Dickens, G.R., 2012. Assessing offsets between the d13C of sedimentary
Soliman, A., Van Nieuwenhove, N., Vink, A., Young, M., 2009. Determining the components and the global exogenic carbon pool across early Paleogene carbon
absolute abundance of dinoagellate cysts in recent marine sediments: The cycle perturbations. Global Biogeochemical Cycles 26, GB4005. http://
Lycopodium marker-grain method put to the test,. Review of Palaeobotany and dx.doi.org/10.1029/2011GB004224.
Palynology 157, 238e252. http://dx.doi.org/10.1016/j.revpalbo.2009.05.004. Thomson, D.J., 1982. Spectrum estimation and harmonic analysis. IEEE Proceedings
Meyers, S.R., Sageman, B., Hinnov, L., 2001. Integrated quantitative stratigraphy of the 70, 1055e1096.
Cenomanian-Turonian Bridge Creek Limestone Member using Evolutive Harmonic Trabucho Alexandre, J., Tuenter, E., Henstra, G.A., van der Zwan, K.J., van de
Analysis and stratigraphic modeling. Journal of Sedimentary Research 71, 627e643. Wal, R.S.W., Dijkstra, H.A., de Boer, P.L., 2010. The mid-Cretaceous North Atlantic
Meyers, S.R., Sageman, B.B., Lyons, T., 2005. Organic carbon burial rate and the nutrient trap: Black shales and OAEs: Paleoceanography v. 25, PA4201.
molybdenum proxy: Theoretical framework and application to Cenomanian- Tsikos, H., Jenkyns, H.C., Walsworth-Bell, B., Petrizzo, M.R., Forster, A., Kolonic, S.,
Turonian OAE II. Paleoceanography v. 20, PA2002. http://dx.doi.org/10.1029/ Erba, E., Premoli Silva, I., Baas, M., Wagner, T., Sinninghe Damste , J.S., 2004.
2004PA001068. Carbon-isotope stratigraphy recorded by the Cenomanian-Turonian oceanic
Meyers, S.R., 2007. Production and preservation of organic matter: The signicance anoxic event: Correlation and implications based on three key localities. Journal
of iron. Paleoceanography 22, PA4211. http://dx.doi.org/10.1029/2006PA001332. of the Geological Society [London] 161, 711e719.
Meyers, S.R., Sageman, B.B., 2007. Quantication of Deep-Time Orbital Forcing by Turgeon, S.C., Creaser, R.A., 2008. Cretaceous oceanic anoxic event 2 triggered by a
Average Spectral Mist. American Journal of Science 307, 773e792. massive magmatic episode. Nature 454, 323e326.
Meyers, S.R., Sageman, B.B., Pagani, M., 2008. Resolving Milankovitch: Consider- Uramoto, G.-I., Tahara, R., Sekiya, T., Hirano, H., 2013. Carbon isotope stratigraphy of
ation of signal and noise. American Journal of Science 308, 770e786. terrestrial organic matter for the Turonian (Upper Cretaceous) in northern Japan:
Meyers, S.R., Hinnov, L.A., 2010. Northern Hemisphere glaciation and the evolution Implications for ocean-atmosphere d13C trends during the mid-Cretaceous cli-
of Plio-Pleistocene climate noise. Paleoceanography 25, PA3207. http:// matic optimum. Geosphere 9, 355e366. http://dx.doi.org/10.1130/GES00835.1.
dx.doi.org/10.1029/2009PA001834. van Helmond, N.A.G.M., Sluijs, A., Reichart, G.-J., Sinninghe Damste , J.S., Slomp, C.P.,
Meyers, S.R., Siewert, S.E., Singer, B.S., Sageman, B.B., Condon, D.J., Obradovich, J.D., Brinkhuis, H., 2014. A perturbed hydrological cycle during Oceanic Anoxic Event
Jicha, B.R., Sawyer, D.A., 2012a. Intercalibration of radioisotopic and astrocho- 2. Geology G34929.1. http://dx.doi.org/10.1130/G34929.1.
nologic time scales for the Cenomanian-Turonian boundary interval, Western Voigt, S., Aurag, A., Leis, F., Kaplan, U., 2007. Late Cenomanian to Middle Turonian
Interior Basin, USA. Geology 40, 7e10. high-resolution carbon isotope stratigraphy: New data from the Mnsterland
Meyers, S.R., Sageman, B.B., Arthur, M.A., 2012b. Obliquity forcing of organic matter Cretaceous Basin, Germany. Earth and Planetary Science Letters 253, 196e210.
accumulation during Oceanic Anoxic Event 2. Paleoceanography 27, PA3212. http://dx.doi.org/10.1016/j.epsl.2006.10.026.
http://dx.doi.org/10.1029/2012PA002286. Voigt, S., Erbacher, J., Mutterlose, J., Weiss, W., Westerhold, T., Wiese, F.,
Meyers, S.R., 2012. Seeing red in cyclic stratigraphy: Spectral noise estimation for Wilmsen, M., Wonik, T., 2008. The Cenomanian-Turonian of the Wunstorf
astrochronology. Paleoceanography 27, PA3228. http://dx.doi.org/10.1029/ section e (North Germany): global stratigraphic reference section and new
2012PA002307. orbital time scale for Oceanic Anoxic Event 2. Newsletters on Stratigraphy 43,
Meyers, S.R., 2014. astrochron: An R Package for Astrochronology. http://cran. 65e89. http://dx.doi.org/10.1127/0078-0421/2008/0043-0065.
rproject.org/packageastrochron. Weedon, G.P., 2003. Time-Series Analysis and Cyclostratigraphy. Cambridge Uni-
Mitchell, R.N., Bice, D.M., Montanari, A., Cleaveland, L.C., Christianson, K.T., versity Press, Cambridge.
Coccioni, R., Hinnov, L.A., 2008. Oceanic anoxic cycles? Orbital prelude to the Westphal, H., Hilgen, F., Munnecke, A., 2010. An assessment of the suitability of
Bonarelli Level (OAE 2). Earth and Planetary Science Letters 267, 1e16. http:// individual rhythmic carbonate successions for astrochronological application.
dx.doi.org/10.1016/j.epsl.2007.11.026. Earth Science Reviews 99, 19e30.
Moore, C.H., 1989. Diagenetic Environments of Porosity Modication and Tools for Wendler, I., 2013. A critical evaluation of carbon isotope stratigraphy and
Their Recognition in the Geologic Record. In: Moore, C.H. (Ed.), Carbonate biostratigraphic implications for Late Cretaceous global correlation. Earth Sci-
Diagenesis and Porosity, Developments in Sedimentology, 46, p. 338. ence reviews 126, 116e146. http://dx.doi.org/10.1016/j.earscirev.2013.08.003.
Pessagno, E.A., 1969. Upper Cretaeous stratigraphy of the western Gulf Coast area of Wilmsen, M., Storm, M., Frsich, F.T., Majidifard, M.R., 2013. Upper Albian and
Mexico, Texas and Arkansas. Geological Society of America memoir 111, 139. Cenomanian (Cretaceous) ammonites from the Debarsu Formation (Yazd Block,
Pratt, L.M., Threlkeld, C.N., 1984. Stratigraphic signicance of 13C/12C ratios in mid- Central Iran). Acta Geologica Polonica 63, 489e513.
Cretaceous rocks of the Western Interior, U.S.A.. In: Stott, D.F., Glass, D.J. (Eds.),
The Mesozoic of Middle North America, 9. Canadian Society of Petroleum Ge-
ologists, pp. 305e312. Appendix A. Supplementary data
Prokoph, A., Villeneuve, M., Agterberg, F.P., Rachold, V., 2001. Geochronology and
calibration of global Milankovitch cyclicity at the Cenomanian-Turonian Supplementary data related to this article can be found at http://dx.doi.org/10.
boundary. Geology 29, 523e526. 1016/j.cretres.2015.04.010.

Você também pode gostar