You are on page 1of 18

See

discussions, stats, and author profiles for this publication at:


http://www.researchgate.net/publication/260633041

Global Milankovitch cycles recorded


in rock magnetism of the shallow
marine lower Cretaceous Cupido
Formation, northeastern Mexico

ARTICLE in GEOLOGICAL SOCIETY LONDON SPECIAL PUBLICATIONS · JULY 2013


Impact Factor: 2.58 · DOI: 10.1144/SP373.20

CITATIONS READS

3 46

5 AUTHORS, INCLUDING:

Linda A. Hinnov
George Mason University
93 PUBLICATIONS 1,601 CITATIONS

SEE PROFILE

Kenneth P. Kodama
Lehigh University
142 PUBLICATIONS 1,755 CITATIONS

SEE PROFILE

Available from: Linda A. Hinnov


Retrieved on: 14 December 2015
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

Geological Society, London, Special Publications

Global Milankovitch cycles recorded in rock magnetism


of the shallow marine lower Cretaceous Cupido
Formation, northeastern Mexico
Linda A. Hinnov, Kenneth P. Kodama, David J. Anastasio, Maya
Elrick and Diana K. Latta

Geological Society, London, Special Publications 2013, v.373;


p325-340.
doi: 10.1144/SP373.20

Email alerting click here to receive free e-mail alerts when


service new articles cite this article
Permission click here to seek permission to re-use all or
request part of this article

Subscribe click here to subscribe to Geological Society,


London, Special Publications or the Lyell
Collection

Notes

© The Geological Society of London 2013


Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

Global Milankovitch cycles recorded in rock magnetism of the shallow


marine lower Cretaceous Cupido Formation, northeastern Mexico
LINDA A. HINNOV1, KENNETH P. KODAMA2*, DAVID J. ANASTASIO2,
MAYA ELRICK3 & DIANA K. LATTA2
1
Department of Earth and Planetary Sciences, Johns Hopkins University,
Baltimore, MD 21218, USA
2
Department of Earth and Environmental Sciences, Lehigh University,
Bethlehem, PA 18015, USA
3
Department of Earth and Planetary Sciences, University of New Mexico,
Albuquerque, NM 87131, USA
*Corresponding author (e-mail: kpkodama@gmail.com)

Abstract: Rock magnetic cyclostratigraphy was measured in the Barremian–Aptian Cupido


(‘Cupidito’) Formation, northeastern Mexico. The goal was to develop an objective evaluation
of palaeo-environmental variability recorded in the formation that is independent of facies analysis
and interpretation. Anhysteretic remanent magnetization (ARM) was used to estimate magnetic
mineral concentration variations for the upper 143 m of the formation, which is characterized by
metre-scale carbonate cycles representative of inner- and middle-shelf marine environments. Iso-
thermal remanent magnetization acquisition experiments and scanning electron microscope (SEM)
examination indicate that micron-sized detrital magnetite from eolian dust carries the ARM signal.
At the sampled sections from Garcia and Chico canyons, 25 km apart, ARM records a synchronous
30–35 m oscillation with maxima coinciding with fourth-order sequence boundaries, superim-
posed with prominent high-frequency variability. Calibrating the 30–35 m oscillation to a
405 kyr period (long eccentricity cycle) focuses the high frequencies into short eccentricity, obli-
quity and precession index bands; the precession-band signal modulates with an eccentricity sig-
nature. The ARM signal is correlated between sections, but decoupled from the interpreted fifth-
order depositional cycles. ARM amplitudes diminish up-section with facies suggesting deepening
conditions that diluted magnetite concentration. This probably signals a warming, increasingly
humid climate, changing global circulation and/or greater dispersal of magnetite grains.

Earth’s ancient shallow marine carbonate platforms change (Ginsburg 1971). To this day, the allocycle
are important archives of past global change. Their v. autocycle debate remains unresolved in under-
in situ biological genesis makes them sensitive indi- standing the origin of 104 –105 year scale lithologic
cators of physical conditions in shallow marine cyclicity in shallow marine carbonates (Schwarza-
environments, including temperature, circulation cher 2000; Burgess 2006; Schlager 2010). Full shal-
and sea level (Wilson 1975). The accommodation lowing to sea level may not develop in response to
space afforded by water depth is thought to govern every sea-level oscillation (subtidal ‘missed beats’),
the amount and type of sediment that accumulates resulting in amalgamated cycles; conversely, pro-
at a given location; water depth in turn is controlled longed episodes of subaerial exposure may occur
by a combination of tectonic subsidence, near-shore in inner platform settings that are only briefly or
deposition and global sea level, which vary signifi- never flooded with every sea-level oscillation,
cantly through time (Schlager 2005). leading to condensed or peritidal ‘missed beats’
The commonly observed stacking of metre- (Goldhammer et al. 1990).
scale shallowing upward cycles in carbonate plat- Ultimately facies analysis of shallow marine sec-
form deposits has led to proposals of allocyclic tions is limited in teasing apart the interplay of
v. autocyclic forcing mechanisms. The most widely eustasy owing to climate change, and subsidence
cited allocyclic hypothesis is astronomically con- arising from local tectonics. This is in large part
trolled eustasy (Fischer 1964; Hardie et al. 1986); owing to the relatively subjective interpretation of
autocyclic mechanisms involve infilling and ocean- lithofacies that is required from the stratigrapher.
ward progradation of sedimentation without eustatic Therefore we need to develop facies-independent

From: Jovane, L., Herrero-Bervera, E., Hinnov, L. A. & Housen, B. A. (eds) 2013. Magnetic Methods and the
Timing of Geological Processes. Geological Society, London, Special Publications, 373, 325– 340.
First published online April 25, 2013, http://dx.doi.org/10.1144/SP373.20 # The Geological Society of London 2013.
Publishing disclaimer: www.geolsoc.org.uk/pub_ethics
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

326 L. A. HINNOV ET AL.

tools to isolate the climate record from sedimen- from the sedimentary facies in which it occurs, but
tary strata. Rock magnetic properties indicate the it is common to both the Garcia and Chico sections.
concentration of magnetic mineral particles and Thus, for this Mexico case study ARM serves the
hence variations in detrital mineral input, for dual purpose of a climate proxy and high-resolution
example, far-field eolian dust influx. Thus, rock correlation tool. The clear implication is that ARM
magnetic parameters have the potential to provide is a transformative method that can elucidate
palaeo-environmental records that are not directly detailed palaeoclimate information from carbonate
linked to facies. stratigraphy with complex origins.
Anhysteretic remanent magnetization (ARM)
has emerged as an important proxy for palaeocli-
mate change in carbonate rocks (Kodama 2012). Geological setting
ARM measures the concentration of ferromagnetic
minerals and is not sensitive to the diamagnetism Regional geology
of the carbonate content; thus it is an ideal tool
for evaluating stratigraphic cyclicity independent Passive margin development along the Gulf Coast
of facies interpretations. Lean & McCave (1998) began in the Late Triassic following continental
demonstrated that glacial –interglacial changes are rifting associated with the opening of the Gulf of
recorded by ARM variations in Pleistocene sedi- Mexico (e.g. Pindell 1993; Dickinson & Lawton
ments on the Chatham Rise, SW Pacific Ocean. In 2001). Uplifted Palaeozoic basement blocks and
those sediments ARM increases during interglacial intervening grabens controlled Mesozoic deposi-
periods in association with decreased productivity, tional patterns along the Mexican Gulf Coast (e.g.
leading to an increased oxic zone in the sediments Wilson 1999; Dickinson & Lawton 2001). In north-
and hence increased production and concentration eastern Mexico, the Coahuila block, presently
of single-domain magnetite grains produced by bounded by the Sabinas and Parras Basins, served
magnetotactic bacteria. Results from Quaternary as a basement high and a proximal source of
marine sediments in the northern Atlantic Ocean clastic material, which influenced evaporite and
show that ARM data can vary on astronomical subsequent carbonate deposition. By the earliest
time scales (Kruiver et al. 1999). ARM has also Cretaceous, extensive carbonate platforms devel-
been used to detect Milankovitch climate cycles in oped around the entire Gulf of Mexico (Wilson
basinal carbonates showing no appreciable facies 1999). In NE Mexico the shelf was shallow and
changes, in deposits coeval to those investigated in rimmed with a stable carbonate platform repre-
the present study (Latta et al. 2006). sented by the late Barremian–Albian Cupido and
Our study focus is on carbonate cycles in the Aurora Formations (Wilson 1999, Lehmann et al.
Lower Cretaceous Cupido Formation as exposed 1999, 2000). The Cupido Formation (940 m thick)
in the Garcia and Chico canyons, northeastern is conformable with respect to the underlying
Mexico (Fig. 1). These cyclic platform deposits shales and limestones of the Barremian to early
have been the subject of numerous facies analysis Aptian Taraises Formation (490 m thick), as well
investigations (Goldhammer et al. 1991; Lehmann as with the overlying the deeper-water shales and
et al. 1998, 1999; Goldhammer 1999; Foster 2003; lime-mudstones of the La Peña Formation (c. 20 –
Latta 2005; Altobi 2007). However, there is no con- 100 m thick).
sensus on the origins of the prominent sedimentary
cyclicity. Stacking patterns suggest an obliquity The Cupido Formation
forcing origin (Goldhammer et al. 1991). Spectral
analysis of multiple sections in the region suggest The Cupido Formation is characterized by stacked,
a combined eccentricity and obliquity forcing ori- metre-scale (1–8 m thick), upward-shallowing peri-
gin (Altobi 2007), but the absence of the precession tidal cycles composed of subtidal facies overlain
index forcing that would be necessary for recording by intertidal to supratidal facies, which in some
of the eccentricity is an unsolved problem. places are capped by subaerial exposure features
Here we have approached this problem using (e.g. karst, palaeosols, rubble breccias; Goldham-
objective rock magnetic data. We measured high- mer et al. 1991; Elrick 1995; Foster 2003). Facies
resolution ARM stratigraphy in the upper Cupido patterns in the upper c. 150 m of the formation
Formation (‘Cupidito’) in Garcia and Chico can- (‘Cupidito’) indicates gradual deepening which is
yons, and characterized the magnetic mineralogy followed by widespread platform drowning rep-
responsible for the ARM variations. The results resented by the La Peña Formation (Wall et al.
led to the discovery of strong astronomical signal in 1961; Tinker 1982; Goldhammer et al. 1991;
the ARM record that can be attributed to climate Lehmann et al. 1998, 1999).
forcing of atmospheric dust deposition to the Our study focuses on two sections on the
region. This ARM signal is largely disconnected Cupido platform: Garcia Canyon, composed of
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 327

Fig. 1. Location and stratigraphic framework of the Lower Cretaceous inner (Garcia) and middle (Chico) carbonate
platform sections in the foreland of the Sierra Madre Oriental fold belt, NE Mexico. (a) Index and hillshade maps
locating the Garcia, Chico and La Boca sections in the foreland of the Sierra Madre Oriental fold belt, NE Mexico.
(b) Regional stratigraphy of northeastern Mexico. The studied section is the upper Cupido (‘Cupidito’) Formation of
latest Barremian– early Aptian age. (c) Cross section through the Sierra del Fraile anticlinorium drawn perpendicular to
the strike of the average fold axes (see a). From Latta & Anastasio (2007).

peritidal-dominated inner-shelf facies; and Chico inner-shelf deposits at Garcia Canyon, Goldhammer
Canyon, dominated by shallow subtidal middle- et al. (1991) identified 78 upward-shallowing cycles
shelf deposits (Fig. 2). The inner- and middle-shelf within a c. 270 m section. Biostratigraphic age
deposits were originally separated by c. 25 km, control indicates a late Barremian to early Aptian
according to cross section restorations of Late Cre- age (Alfonso-Zwanzinger 1978). Goldhammer et al.
taceous Sevier– Laramide deformation (Latta & (1991) interpreted the cycles (average thickness
Anastasio 2007). The inner-shelf deposits are 3.44 m) to represent 40 kyr obliquity forcing based
characterized by thin-bedded, peritidal and subtidal on age, number of cycles and cycle bundling ratios
facies, including evaporite, laminated mudstone, between sequence boundaries.
and burrowed skeletal packstone units. The middle- The inner- and middle-shelf deposits of the
shelf deposits contain similar peritidal and subtidal upper Cupido Formation at Garcia and Chico Can-
facies, but exhibit less frequent (less cyclic) facies yons are overlain by the relatively isochronous
changes, resulting in thicker-bedded units. In the basal La Peña Formation (Goldhammer 1999;
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

328 L. A. HINNOV ET AL.

NE inner middle outer SE


shelf 25 km
shelf shelf sea level

shelf crest

20 m
shoal complex shelf edge
rudist reef
La Peña Formation
AP4 datum (121.0 myr)

Peritidal Shallow Subtidal

Cupido Formation
Cupido Formation

evaporites skeletal
packstone

143 m
143 m

laminites skeletal
mud/grainstone
heterolithic non-skeletal
thin beds grainstone

AP3 datum (123.97 myr)

AP2 datum (124.58 myr)


Garcia Chico

Fig. 2. Coeval inner- and middle-shelf deposits of the upper Cupido (‘Cupidito’) Formation. Regionally correlative
datums (AP2, AP3 and AP4, see text) provide chronostratigraphic constraints, and exposure facies allow
lithostratigraphic correlation. Pie charts illustrate facies distribution across the platform and indicate that the inner-shelf
deposits at Garcia are dominated by peritidal facies and the middle-shelf deposits at Chico are dominated by
subtidal facies.

Lehrmann & Goldhammer 1999; Fig. 2). The bio- and from 117.07 to 121.0 Ma, respectively (Har-
stratigraphic controls on the La Peña Formation are denbol et al. 1998; updated by Snedden & Liu
ambiguous. At Santa Rosa Canyon (SW of the 2010). These sequence boundaries can be corre-
study area), the La Peña starts in the upper Globi- lated to biostratigraphically constrained sequence
gerinelloides blowi planktonic foraminifer zone boundaries in the northern Gulf of Mexico (Gold-
(Bralower et al. 1999), which places it below Ocea- hammer 1999), guyots in the Western Pacific
nic Anoxic Event 1a (OAE1a, the ‘Selli’ event) (Röhl & Ogg 1998) and western European sedimen-
described by Schlanger & Jenkyns (1976), Coccioni tary basins (Jacquin et al. 1998). Thus these datums
et al. (1987, 1989, 2012), and many others. How- appear to represent global sea level changes.
ever, in La Huasteca Canyon (south of the study However, new astrochronology of the Italian
area), the La Peña starts in the Dufrenoyia furcata Piobbico core indicates that OAE1a starts at
(¼Dufrenoyia justinae Zone in Mexico) ammonite 124.55 Ma, and ends at 123.16 Ma (Huang et al.
zone (Barragan & Maurrasse 2008), which places 2010). Thus, if the basal La Peña pre-dates OAE1a
it above OAE1a. (i.e. is older than 124.55 Ma, as indicated by fora-
The Garcia and Chico sections have three minifera stratigraphy) then these datums may need
chronostratigraphically significant sequence bound- to be reassigned to older sequence boundaries, for
aries in the upper Cupido Formation (Fig. 3). example, Barr6 (126.2 Ma), AP1 (125.0 Ma) and
These regionally correlative datums were originally AP2 (124.58 Ma). On the other hand, if the basal
assumed to correspond to global sequence bound- La Peña post-dates OAE1a (i.e. is younger than
aries for AP2, AP3 and AP4 (Goldhammer 1999), 123.16 Ma, as indicated by ammonite stratigraphy),
for which dates have been recently updated from then the presently assumed global sequence bound-
120.6 to 124.58 Ma, from 120.0 to 123.97 Ma ary assignments may stand.
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 329

2 additional sequence boundaries


AP2 AP3 (Goldhammer, 1999) AP4
20
0

feet -20
-40
-60
-80
-100
10 20 30 40 50
cycle no.
Fig. 3. Fischer plot of the Cupidito carbonate cycle succession at Garcia Canyon. Arrows indicate sequence boundaries
as interpreted by Goldhammer (1999). The section duration is 3.58 or 1.62 myr (see text), which in turn indicates a
cycle duration of 73 or 33 kyr, and sequence durations of c. 730 –900 or 330– 407 kyrs. There are 10–12 cycles per
sequence, that is, no sustained 5:1 cycle bundling.

Data and methods resulted in a total of 297 samples at Garcia and 283
samples at Chico.
Sample collection
Samples were collected from the upper Cupido For- Rock magnetic analysis
mation at Garcia (inner-shelf) and Chico (middle-
shelf) canyons. The sample spacing was maintained Individual, unorientated samples were collected and
between 20 and 50 cm in order to sample each iden- hand crushed to 2– 4 mm size pieces and weighed
tified carbonate cycle at least four or five times. This in individual, pre-weighed, 8 cm3 plastic sample

Fig. 4. Representative low-temperature magnetic measurements for inner (a) and middle (b), shelf deposits, indicating
a sharp transition (c. 100 K) upon cooling (black circles), suggesting Verwey transitions, and that the ARM signal is
probably controlled by ferromagnetic magnetite (e.g. Muxworthy 1999). From Latta (2005).
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

330 L. A. HINNOV ET AL.

boxes. Low-field bulk magnetic susceptibility was and the magnetic moment of each sample was again
measured on a Agico KLY-3s Kappabridge, while measured in a 0 T field at 5 K increments while
ARM was acquired in a peak alternating field of cooled back to 20 K.
100 mT with a DC field of 0.1 mT and measured on The relative populations of magnetic grains,
a 2G Enterprises, Inc. superconducting magneto- based on variations in grain size and the mineralogy
meter. Both measurements were normalized for of the ferromagnetic grains, were determined from
mass and all were conducted at Lehigh University. 32 samples (16 from inner-shelf Garcia, and 16
Variations in ARM can be caused by fluctuations from middle-shelf Chico) using the vibrating sam-
in primary ferromagnetic mineral concentrations ple magnetometer (VSM1) where the field varied
or from diagenetic processes. In order to further between 1.0 and –1.0 T. In an additional experiment
characterize magnetic grain size and composition, to determine the magnetic mineralogy, isothermal
rock magnetic experiments were undertaken at the remanences (IRMs; Lowrie 1990) were imparted
Institute for Rock Magnetism, University of Minne- to a single outer-shelf sample from La Boca Can-
sota. Ferromagnetic and paramagnetic grains were yon that displayed relatively high ferromagnetic
characterized in 24 samples (11 from the inner-shelf mineral concentrations and were subsequently ther-
Garcia section and 13 from the middle-shelf Chico mally demagnetized (Latta et al. 2006). Finally,
section) using the superconducting susceptometer scanning electron microscope/energy dispersive
(MPMS). After applying a 2.5 T field, the magnetic X-ray microscopy images of magnetic grains col-
moment of each sample was measured in a 0 T lected from HCl-insoluble residues of samples that
field at 5 K increments while heated from 20 to display hysteresis parameters characteristic of all
300 K. A 2.5 T field was then again applied at 300 K samples, were used to help elucidate ferromagnetic
grain morphology and size.

Time series methods


To study the frequency content of the ARM records
we used the multitaper spectral analysis function in
freeware Analyseries (Thomson 1982; Paillard et al.
1996). The ARM series were resampled to uniform
spacing using the linear interpolation in Analyseries.
The Garcia facies rank series was interpolated to
a uniform spacing using the ‘nearest neighbor’
option in Matlab’s interp1 function. Taner bandpass
filtering (Taner 2003) was applied in Matlab to
isolate the precession index signal in the 405 kyr
tuned Garcia ARM record. Amplitude envelope
analysis was performed in Matlab using Hilbert
transformation (Hinnov 2000).

Results
Magnetic mineralogy
The MPMS results indicate the presence of magne-
tite as the dominant magnetic mineralogy in the
Cupido Formation carbonates. The solid circles in
Fig. 5. Experimental results showing hysteresis curves Figure 4 indicate the cooling cycle in the MPMS
of representative samples from each measured section, measurements. The abrupt decrease in magnetic
with the diamagnetic component removed. intensity at about 100 K in the cooling curve is inter-
(a) Inner-shelf deposits exhibit two different grain size preted to be the Verwey transition of magnetite. The
distributions of the magnetic mineral, most likely presence of magnetite is important since it indicates
magnetite based on the low-temperature results, as a primary depositional magnetic mineral carrying
shown by the wasp-waisted hysteresis curves, indicating
a mixture of single domain and superparamagnetic the rock magnetic cyclostratigraphy in the Cupido
grains. (b) Middle-shelf deposits, indicates the presence Formation.
of a single low-coercivity ferromagnetic mineral The hysteresis curves (Fig. 5) for the two
population, probably in the pseudo-single domain grain sampled localities show a wasp-waisted loop for
size range. the inner shelf (Garcia) and a narrow loop for
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 331

0.7 (Fig. 6) suggests that the magnetic particle size


for the magnetite in the Cupido rocks is the vortex
0.6 magnetic state (Tauxe et al. 2002), suggesting
pseudo single domain grains with a grain size of
0.5
c. 0.115 mm, with slightly larger particles in the
outer shelf (Chico). Finally, anhysteretic remanent
magnetization/saturation isothermal remanence
0.4
magnetization values of 0.05–0.2 and Raf values
Mr/Ms

of 0.35 –0.4, where Raf is the crossover point


0.3 Chico between the IRM acquisition curve and the af
Garcia
demagnetization curve of the SIRM, suggest that
0.2 vortex no magnetosomes are present in the Cupido car-
bonates. Magnetosomes are produced by magneto-
0.115
0.1 tactic bacteria (Bazylinski et al. 1988).
The rock magnetic results, therefore, indicate
MD that the dominant magnetic mineral carrying the
0
0 10 20 30 40 50 60 rock magnetic cyclostratigraphy is primary, deposi-
Hc (mT) tional inorganic magnetite grains. SEM/EDX
reveals numerous iron oxide grains ranging in size
Fig. 6. Magnetic granulometry at Garcia and Chico from 1.5 to 12 mm, where nearly 50% of the
canyons. This squareness diagram indicates that most of grains fall between 3.0 and 4.0 mm (Fig. 7). SEM
the samples measured from Garcia (red circles) and
analyses further support the interpretation that the
Chico (blue squares) are in the vortex to MD state,
suggesting grain sizes of 0.1– 1 mm. ferromagnetic grains are detrital. The iron oxide
grains are observed to have quartz coatings, as
expected from weathered bedrock, and an absence
the middle shelf (Chico). The wasp-waisted loop of textures consistent with diagenetic processes.
indicates a mixture of single domain and superpara-
magnetic ferromagnetic grains in the rocks (Tauxe ARM cyclostratigraphy
et al. 1996), suggesting a range of magnetic particle
sizes in the sub-micron range. The narrow hystere- The raw ARM measurements show a strong
sis loop from the middle shelf indicates pseudo- decrease in magnitude upsection (Figs 8 & 9).
single domain grains. The squareness (Mr/Ms) This trend corroborates facies observations of
v. coercivity (Hc) plot of the hysteresis parameters increasing water depths (Fig. 2) and flooding of the

Fig. 7. Magnetite grain sizes analysed from HCl-insoluble residues by SEM/EDX. Average grain size 3.3 mm
(+1 mm) is consistent with that of far-travelled atmospheric dust particles. Inset background electron image is of an
iron-oxide grain from the Garcia section with quartz coatings, suggestive of original quartz cement and detrital origin.
From Latta (2005).
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

332 L. A. HINNOV ET AL.

Fig. 8. (a) Comparison of ARM values between inner- (Garcia) and middle- (Chico) shelf measured sections in the
upper Cupido Formation. The length of the sampled Garcia section is 136 m (7– 143 m), and for the Chico section is
131 m (6 –137 m). Up is to the left. (b) Detailed view of the upper 60 m; note the close correlation of ARM values
marked by representative peaks (A –I) From Latta (2005).

Cupido Platform and subsequent deposition of the in both sections (Fig. 9). The fitted linear trends
deep marine La Peña Formation. The Garcia and show that ARM declines at the same rate in both
Chico records share numerous ARM peaks in strati- sections; the offset of the trends shows that Garcia
graphic order (labelled A –I in Fig. 8b), which attest has a higher overall ARM throughout, presumably
to a remarkable consistency in magnetic mineral owing to its shallower depositional site. Most spec-
input between the two localities, which were sepa- tacularly the sections share a strong, phase-locked
rated by c. 25 km and had generally different c. 30 m cycle. This cycle is also captured in the
water depths. Fischer plot of the Garcia section (Fig. 3). This
Log transforming the ARM (‘log10ARM’) dra- Fischer plot was used by Goldhammer (1999) to
matically improves our view of the ARM signal, interpret fourth-order sequence boundaries, which
revealing strong ARM cyclicity at multiple scales coincide with the ARM maxima of the c. 30 m
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 333

Garcia stratigraphic position (m)


140 120 100 80 60 40 20
143 m 122 m 88 m 58 m 32 m
-4.4
4th order sequence boundaries

-4.8 INNER SHELF Garcia


MIDDLE SHELF Chico

-5.2
log10(ARM)

-5.6

-6

-6.4

-6.8 4th order sequence boundaries


140 m 118 m 85 m 56 m 33 m

120 100 80 60 40 20
Chico stratigraphic position (m)

Fig. 9. Log10 transformed ARM record for Garcia (red) and Chico (blue). Up is to the right. Fourth-order sequence
boundaries are from Goldhammer (1999). The linear trend shows identical decline in ARM in both sections because of
the deepening depositional environment upsection. The Garcia trend is offset from Chico owing to Garcia’s deposition
in a shallower setting. The cross-correlation coefficient between the two records, assuming that the bases and tops (basal
contact of the La Peña Formation) of the sections are time-correlative points, is 0.68.

cycle (Fig. 9). Finally, superimposed on this domi- et al. 1991; Goldhammer 1999; Foster 2003; Latta
nant cycle are high-frequency variations that 2005) that recognize pronounced cyclicity and
persist through both sections. intervals of subaerial exposure, thus providing evi-
dence for sea-level oscillations. Therefore, in this
discussion we focus on analysing the ARM record
Spectral analysis
at Garcia, seeking evidence for astronomical
Power spectral analysis of the Garcia and Chico forcing, and comparing it with the upward shallow-
log10ARM records in the stratigraphic domain ing carbonate cycles.
(Fig. 10) confirm that the two records share impor-
tant spectral characteristics. The 33.3 m cycle
dominates both spectra; the Chico spectrum has ARM as eolian dust proxy
high-power peaks at 14.9 and 11.4 m, whereas at The SEM/EDX analysis reveals that iron oxide
Garcia there is only one, low-power peak at 13.5 m. shapes probably represent detrital grains (Fig. 7).
Power declines rapidly in both spectra at higher This, together with the small sizes of the analysed
frequencies, where spectral peaks occur at differ- grains, is consistent with a far-travelled atmo-
ent frequencies. These mismatched peaks are due spheric dust origin (e.g. Pye 1987). Early Cretac-
to different sedimentation rates affecting the two eous atmospheric circulation indicates easterly
depositional environments. The Garcia spectrum, trade winds affecting ancestral Mexico (e.g. Hase-
with its greater loss of power at higher frequencies, gawa et al. 2012). This could allow transport of
may reflect some abbreviation (condensation) of eolian material from ancestral Africa across the
the depositional record owing to the higher inci- widening Atlantic Ocean (Fig. 11), similar to what
dence of subaerial exposure. occurs today (Goudie & Middleton 2006). In
support of our ARM evidence in Mexico, Deep
Discussion Sea Drilling Project Sites 387, 391C, and 367
from the North Atlantic Basin contain Lower Cre-
Multiple sedimentological investigations have been taceous eolian dust transported from northwestern
undertaken on the Garcia section (Goldhammer Africa (Dean & Arthur 1999). This eolian flux to
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

334 L. A. HINNOV ET AL.

(a) 16
33.3 m
14
12
10
power

8 13.5 m
6 9.2 m
4
3.77 m
2 1.7 m
0
0 0.2 0.4 0.6 0.8 1
cycles/m
(b) 5
33.3 m
4 14.9 m
11.4 m
3
7.2 m
2
2.0 m
3.2 m
1

0
0 0.2 0.4 0.6 0.8 1
cycles/m
Fig. 10. 2p Power spectrum of the Garcia log10ARM (a) and Chico log10ARM (b) records in the stratigraphic domain.
Labels indicate peak cyclicities in metres.

Fig. 11. Lower Cretaceous (115 Ma) palaeogeography, modified from Scotese et al. (1989), showing possible routes of
eolian dust from Africa to the depositional site in Mexico (black star). Dark grey, continents; light grey, shallow shelves;
white, ocean. From Latta et al. (2006).
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 335

Mexico would therefore be expected to reflect cli- AP2, then the duration is 1.62 myr. Both of these
matic forcing that affected low latitude atmospheric estimated durations have unknown certainties that
circulation. At the 104 –105 year timescales resolved are probably in the half-million year range based
by the Cupido carbonate cycles, evidence for on radioisotope dating constraints (Ogg & Hinnov
astronomical forcing of eolian processes should be 2012).
in sharp focus. We tested the hypothesis that the 33.3 m
log10ARM cycles were caused by 405 kyr orbital
eccentricity forcing by tuning the former to the
The ARM astronomical signal at Garcia latter and assessing periodicity in the tuned spec-
Canyon trum. This timing corresponds to the alternative
Barr6-AP2 assignment discussed above. The results
The Garcia log10ARM spectral peaks (Fig. 10a) (Fig. 12) show focusing of power in the 100, 44 and
have a frequency distribution that is suggestive of 20 kyr bands, as might be expected for an astro-
astronomical forcing (33.3:13.5:9.2:3.77:1.7 m is nomical signal. Strong forcing by the orbital eccen-
close in ratio to 405:130:97:c. 40:20 kyr). There is tricity, however, is a misnomer. The actual source
elevated power in the frequency band correspond- of eccentricity power in stratigraphy is from pre-
ing to the precession index (1.7 m), but strong cession index forcing. The stratigraphic recording
peaks do not occur, possibly owing to variable sedi- process transfers power from precession index
mentation rates changing the thickness of preces- cycling into the precession modulator, that is, the
sion scale cycles through the section. eccentricity, in an effect known as ‘rectification’
At present, there is not an accurate chronology (Weedon 2003). To examine whether there is a
for these sections. Biostratigraphic evidence cannot direct connection between the recorded preces-
establish whether the top of the Cupidito (basal La sion index and eccentricity, we filtered the preces-
Peña) is pre-OAE1a or post-OAE1a in age; no geo- sion band (see Data and Methods section) and
chronology is directly tied to either section. The compared the amplitude envelope of the filtered
global sequences AP2, AP3 and AP4 interpreted signal with the eccentricity as recorded in the
by Goldhammer (1999) (Figs 2 & 3) indicate a dura- original tuned log10ARM signal. Filtered signal
tion of 3.58 myr for the section. Alternatively, if amplitudes are high when 405 kyr cycles reach
these global sequences are instead Barr6, AP1 and their maxima, and low when the cycles are at

(a) -4 (b) 20
405-kyr
-4.5
-5 15
0.035, 0.045, 0.055
logARM

precession
power

-5.5 passband
10
-6
100 kyr
-6.5
5
-7 44 kyr 20 kyr
-7.5 0
-500 0 500 1000 1500 2000 0 0.02 0.04 0.06 0.08 0.1
kiloyears upsection cycles/kyr
(c) (d) 1
0.4
LogARM (residual)

0.8 400 kyr


0.2
0.6
power

0
0.4
-0.2 95 kyr
0.2
-0.4
0
-500 0 500 1000 1500 2000 0 0.01 0.02 0.03 0.04 0.05
kiloyears upsection cycles/kyr

Fig. 12. Garcia log10ARM record calibrated to 405 kyr cycles. (a) 405 kyr calibrated Garcia logARM time series. The
405 kyr intervals were defined at 122, 88, 58 and 32 m (sequence boundaries of Goldhammer 1999) in the Garcia
section. The total length of this time series is 1.84556 myr. (b) 2p Power spectrum of the 405 ky-calibrated Garcia
log10ARM time series. The passband of the Taner filter (Taner 2003) applied to extract a precession index signal is
shown, where the cut-off frequencies are 0.035 and 0.055 cycles kyr21, and the centre frequency of the band is
0.045 cycles kyr21. (c) Taner-filtered precession index signal (red line, grey in print version; passband shown in b) and
amplitude envelope (black line) obtained through Hilbert transformation (Hinnov 2000). (d) 2p Power spectrum of the
amplitude envelope shown in (c).
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

336 L. A. HINNOV ET AL.

their minima (compare Fig. 12a, c), supporting the recorded as water depths are greater, resulting in
hypothesis that the c. 1.7 m scale ARM cycles were greater dispersion of eolian materials and dilution
precession-driven. by increased carbonate accumulation.
However, the individual carbonate cycles are not
ARM record v. depositional cyclicity matched by any of the high-frequency ARM cycles.
Instead, there are two or more ARM cycles per car-
Comparison of the ARM signal and carbonate bonate cycle (Fig. 13b). Thus there is a decoupling
cycles at Garcia reveals some surprises. The max- between the sea-level oscillations forcing carbonate
ima of the ARM 33.3 m cycles correlate with thin- cycle development and the eolian influx producing
ner and more subaerially exposed intervals (Fig. the ARM signal. The strong astronomical signature
13a). This suggests that, when sea level is low, high in the ARM signal is not matched by a comparable
ARM associated with increased windiness and one in carbonate cycling. The power spectra of the
atmospheric dustiness is recorded in the Cupido two Garcia records (Fig. 14) show that they share
Platform. When sea level is high, low ARM is a common 405 kyr cycle, but at higher frequencies,

Fig. 13. Decoupling of the log10ARM record v. depositional cycles at Garcia Canyon. (a) The entire section.
(b) Close-up of 70–140 m, where the log10ARM record has been prewhitened to remove the linear trend and 33 m cycle.
Lithologic classes: 1, subaqueous gypsum; 2, laminites; 3, heterolithic thin beds; 4, thalassanoides burrowed
wackestones; 5, non-skeletal grainstones; 6, bedded requienids and chrondrodonts; 7, skeletal packstones; 8, rudistid
bioherms. (Adapted from Foster 2003.)
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 337

(a) 20
405-kyr

15
power
10
100 kyr
5
44 kyr 20 kyr
0
0 0.02 0.04 0.06 0.08 0.1
cycles/kyr
(b) 1
405-kyr
0.8
180 kyr
72 kyr
0.6
power

58 kyr
0.4

0.2

0
0 0.02 0.04 0.06 0.08 0.1
cycles/kyr
Fig. 14. 2p Power spectra of the 405 kyr calibrated ARM time series (a) and facies rank series (b) shown in Figure 13a.

where the ARM displays astronomical frequencies 110 Ma. Later, the sequences were correlated into
(spectral peaks at periods of 100, c. 44 and the Hardenbol et al. (1998) global sequence frame-
20 kyr), the facies rank series spectrum has strong work with AP2 at the base of the Cupidito and AP4
spectral peaks at periods of 72 and 58 kyr, neither at the top (Lehmann et al. 2000). AP3 is assigned
of which is recognizable as an astronomical period. to the sequence boundary within the Cupidito
However, since the two records share a strong associated with the most intense features of sub-
405 kyr cycle, the 72 and 58 kyr periods in the aerial exposure (Fig. 2). This framework, however,
facies spectrum may caused by of ‘missed beats’ does not account for two additional fourth-order
in response to sea-level oscillations originating sequences within the Cupidito indicated by Gold-
from high latitude astronomically forced ice sheets. hammer (1999) (Fig. 3). These sequence boundaries
coincide with ARM maxima of the 33.3 m cycles,
Cupidito sequence stratigraphy except at 32 m (Garcia) and 33 m (Chico) (Fig. 9).
The evidence presented above for a 405 kyr
The fourth-order sequences (Fig. 3) are part of a (orbital eccentricity) timing of the Goldhammer
‘supersequence’ encompassing the entire Cupidito fourth-order sequences suggests that the Cupidito
section that progresses from a lowstand to a trans- section is 1.846 myr in duration (Fig. 12). This dur-
gressive systems tract (Goldhammer 1999). This ation is close to the original time frame indicated by
designation accounts for deepening trends in the Goldhammer (1999), and not the longer chronology
upper Cupido Formation (‘Cupidito’) and the La (3 myr) of the Hardenbol et al. (1998) global
Peña drowning. The base of the Cupidito (the low- sequence framework, or the updated chronology
stand) was originally assigned an age of 112 Ma; (3.58 myr) by Snedden & Liu (2010). The ambigu-
the top of the Cupidito was assigned an age of ity imposed by biostratigraphy (La Peña either
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

338 L. A. HINNOV ET AL.

pre- or post-OAE1a) allows that the top of the stored in the carbonates is an ‘atmospheric’ proxy.
Cupidito may be pre-OAE1a, which could indicate Both are far-field proxies: sea-level changes origi-
a AP2 global sequence boundary with an age of nated from the dynamics of continental ice sheets
124.58 Ma. Proceeding downward for 1.846 myr and other water reservoirs, as well as from local tec-
to 126.42 Ma brings us close to the Barr6 sequence tonics, and eolian dust flux was controlled by
boundary age of 126.2 Ma. However, this still changes in continental aridity and global wind pat-
leaves us with two extra sequences that are not terns. We anticipate that, in the future, the develop-
described in the global sequence framework. To ment of similar multiple proxies in shallow marine
solve this problem we need new chronostratigra- carbonate sections such as the Cupido Forma-
phic data. Specifically, carbon isotope stratigraphy, tion will transform our understanding of the local
which has a distinctive evolution through the depositional conditions as well as the processes
OAE1a interval (Li et al. 2008; Malinverno et al. that controlled deposition. Rock magnetic pro-
2010), may help determine whether the Cupidito perties of carbonates in particular provide the
carbonates are pre- or post-OAE1a in age. additional, very powerful, proxies for such studies.
L. H. was supported by US National Science Founda-
tion grant EAR-0718905. We thank M. Jackson and
Conclusions P. Solheid for analytical assistance and advice during
D. L.’s fellowship at the Institute of Rock Magnetism, Uni-
ARM of the upper Cupido (‘Cupidito’) Formation, versity of Minnesota. K. K., D. A. and M. E. were supported
northeastern Mexico, was measured at two key sec- by US National Science Foundation grant EAR-0230053;
tions at Garcia and Chico canyons to investigate the D.L. was also supported by a student grant from the Geo-
origin of the metre-scale carbonate cyclicity. The logical Society of America. This manuscript is an extension
of D. L.’s PhD dissertation at Lehigh University.
results uncovered an unexpectedly strong astronom-
ical signal hidden in the ARM stratigraphy. This
signal is common to both of the studied sections,
which were separated by 25 km, and accumulated References
in different water depth conditions. ARM proved Alfonso-Zwanzinger, J. 1978. Geologia regional de
to be a powerful tool for correlating these carbonate sistema sedimentario Cupido. Boletin de la Asociacion
sections at extremely high resolution (within a few Mexicanna de Geologos Petroleros, 30, 1– 56.
thousand years). Altobi, Y. K. 2007. Milankovitch orbital forcing control
The source of the ARM in the carbonates was on shallow-water carbonate cyclicity and early dolo-
traced to fine-grained detrital magnetite consistent mitization: insights from the Lower Cretaceous
with eolian dust. Early Cretaceous models of atmos- Cupido Platform, NE Mexico. PhD dissertation, Uni-
pheric circulation indicates that the most probable versity of Texas at Austin, Austin, TX.
source of this dust would have been from northeast- Barragan, R. & Maurrasse, F. J.-M. R. 2008. Lower
Aptian (Lower Cretaceous) ammonites from the La
ern Africa; easterly trade winds would have trans- Peña Formation of Nuevo Leon State, northeast
ported the dust across the North Atlantic Basin to Mexico: chronostratigraphic implications. Revista
the depositional site in Mexico. Astronomical Mexicana de Ciencias Geologicas, 25, 145– 157.
forcing controlled the intensity and/or direction of Bralower, T. J., CoBabe, E., Clement, B., Sliter, W. V.,
the winds, and/or aridity on the African continent, Osburn, C. L. & Longoria, J. 1999. The Record
modulating the amount of dust arriving at the of global change in mid-Cretaceous (Barremian-
Cupido Platform. Albian) sections from the Sierra Madre, northeastern
While the astronomical signal in the Cupidito Mexico. Journal of Foraminiferal Research, 29,
418–437.
ARM record is strong, the cyclic carbonate stratigra-
Burgess, P. M. 2006. The signal and the noise: forward
phy lacks a clear astronomical signal, although modeling of allocyclic and autocyclic processes influ-
it has some attributes associated with 405 kyr encing peritidal carbonate stacking patterns. Journal
orbital eccentricity forcing. This corroborates long- of Sedimentary Research, 76, 962– 977.
standing suspicions that shallow marine carbonates Coccioni, R., Nesci, O., Tramontana, M., Wezel, F. C.
are imperfect recorders of astronomical forcing; & Moretti, E. 1987. Descrizione di un livello–guida
at the same time, the rock magnetism approach ‘radiolaritico– bituminoso– ittiolitico’ alla base delle
offers a solution to this problem. Since the platform Marne a Fucoidi nell’Appennino Umbro– Marchi-
recorded sea-level variations that had at least a pro- giano. Bollettino della Società Geologica Italiana,
minent 405 kyr cycling at Garcia, we can infer that 106, 183 –192.
Coccioni, R., Franchi, R., Nesci, O., Wezel, F. C., Bat-
the sea-level oscillations had attributes of insola- tistini, F. & Pallecchi, P. 1989. Stratigraphy and
tion forcing that included a strong precessional mineralogy of the Selli Level (Early Aptian) at the
component. base of the Marne a Fucoidi in the Umbro-Marchean
In sum, in this study the carbonate facies cycli- Apennines, Italy. In: Wiedmann, J. (ed.) Cretaceous
city is an ‘oceanic’ proxy, and the eolian dust of the Western Tethys. Proceedings 3rd International
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

GLOBAL MILANKOVITCH, CRETACEOUS CUPIDO 339

Cretaceous Symposium. E. Schweizerbart’sche Ver- high frequency sea level oscillations on a 104 year
lagsbuchhandlung, Stuttgart, 563– 584. scale. Paleoceanography, 1, 447 –457.
Coccioni, R., Jovane, L. et al. 2012. Umbria-Marche Hasegawa, H., Tada, R. et al. 2012. Drastic shrinking of
Basin, Central Italy: a reference section for the the Hadley circulation during the mid-Cretaceous
Aptian-Albian interval at low latitudes. Scientific Dril- Supergreenhouse. Climates of the Past, 8, 1323– 1337.
ling, 13, 42– 46, doi:10.2204/iodp.sd.13.07.2011. Hinnov, L. A. 2000. New perspectives on orbitally forced
Dean, W. E. & Arthur, M. A. 1999. Sensitivity of the stratigraphy. Annual Review of Earth and Planetary
North Atlantic basin to cyclic climatic forcing during Sciences, 28, 419–475.
the Early Cretaceous. Journal of Foraminiferal Huang, C., Hinnov, L. A., Fischer, A. G., Grippo, A. &
Research, 29, 465– 486. Herbert, T. 2010. Astronomical tuning of the Aptian
Dickinson, W. R. & Lawton, T. F. 2001. Carboniferous stage from Italian reference sections. Geology, 238,
to Cretaceous assembly and fragmentation of Mex- 899– 903.
ico. Geological Society of America Bulletin, 113, Jacquin, T., Rusciadelli, G., Amedro, F., de Gra-
1142–1160. ciansky, P.-C. & Magniez-Jannin, F. 1998. The
Elrick, M. 1995. Cyclostratigraphy and sequence strati- North Atlantic cycle; an overview of 2nd-order trans-
graphy of the Lower Cretaceous Cupido Formation, gressive/regressive facies cycles in the Lower Cretac-
northeastern Mexico. Annual Meeting Abstracts – eous of Western Europe. In: de Graciansky, P.-C.,
American Association of Petroleum Geologists and Hardenbol, J., Jacquin, T. & Vail, P. R. (eds) Meso-
Society of Economic Paleontologists and Mineralo- zoic and Cenozoic Sequence Stratigraphy of European
gists, 4, 26. Basins. Society for Sedimentary Geology, Tulsa, OK,
Fischer, A. G. 1964. The Lofer cyclothems of the Alpine Special Publications, 60, 397– 409.
Triassic. In: Merriam, D. F. (ed.) Symposium on Kodama, K. P. 2012. Paleomagnetism of Sediments and
Cyclic Sedimentation, Lawrence, KS. Kansas Geologi- Sedimentary Rocks: Process and Interpretation.
cal Survey Bulletin, 169, 107–149. Wiley, Chichester.
Foster, T. R. 2003. The evolution of a Lower Cretaceous Kruiver, P. P., Kok, Y. S., Dekkers, M. J., Langereis, C.
carbonate platform within a divergent margin setting: G. & Laj, C. 1999. A pseudo-Thellier relative paleoin-
The Cupido Formation, northeastern Mexico. MSc tensity record, and rock magnetic and geochemical par-
thesis, The University of Texas at Austin, Austin, TX. ameters in relation to climate during the last 276 kyr in
Ginsburg, R. N. 1971. Landward movement of carbonate the Azores region. Geophysical Journal International,
mud: a new model for regressive cycles in carbonates 136, 757–770.
(abstract). American Association of Petroleum Geolog- Latta, D. K. 2005. Structural, Lithotectonic, and Rock
ists, Annual Conference, Program Abstracts, 55, 340. Magnetic Studies of Decollement Folding, Coahuila
Goldhammer, R. K. 1999. Mesozoic sequence stratigra- Marginal Folded Province, NE Mexico. PhD disser-
phy and paleogeographic evolution of northeast Mex- tation, Lehigh University, Bethlehem, PA.
ico. In: Bartolini, C., Wilson, J. L. & Lawton, T. Latta, D. K. & Anastasio, D. J. 2007. Multiple scales of
F. (eds) Mesozoic Sedimentary and Tectonic History mechanical stratification and decollement fold kine-
of North-Central Mexico. Geological Society of matics, Sierra Madre Oriental foreland, northeast Mex-
America, Boulder, CO, Special Papers, 340, 1 –58. ico. Journal of Structural Geology, 29, 1241–1255.
Goldhammer, R. K., Dunn, P. A. & Hardie, L. A. 1990. Latta, D. K., Anastasio, D. J., Hinnov, L. A., Elrick,
Depositional cycles, composite sea-level changes, M. & Kodama, K. P. 2006. A magnetic record of
cyclic stacking patterns, and the hierarchy of strati- Milankovitch rhythms in lithologically non-cyclic
graphic forcing: examples from Alpine Triassic plat- marine carbonates. Geology, 34, 29– 32.
form carbonates. Geological Society of America Lean, C. M. B. & McCave, I. N. 1998. Glacial to intergla-
Bulletin, 102, 535– 562. cial mineral magnetic and palaeoceanographic changes
Goldhammer, R. K., Lehmann, P. J., Todd, R. G., at Chatham Rise, SW Pacific Ocean. Earth and Plane-
Wilson, J. L., Ward, W. C. & Johnson, C. R. 1991. tary Science Letters, 163, 247– 260.
Sequence Stratigraphy and Cyclostratigraphy of the Lehmann, C., Osleger, D. A. & Montanez, I. P. 1998.
Mesozoic of the Sierra Madre Oriental, Northeast Controls on cyclostratigraphy of Lower Cretaceous
Mexico, a Field Guidebook. Gulf Coast Section, carbonates and evaporates, Cupido and Coahuila plat-
Society of Economic Paleontologists and Mineralo- forms, northeastern Mexico. Journal of Sedimentary
gists, Tulsa, OK. Research, 68, 1109– 1130.
Goudie, A. S. & Middleton, N. J. 2006. Desert Dust in Lehmann, C., Osleger, D. A., Montañez, I. P., Sliter,
the Global System. Springer, Berlin. W., Arnaud-Vanneau, A. & Banner, J. 1999. Evol-
Hardenbol, J., Thierry, J., Farley, M. B., Jacquin, T., ution of Cupido and Coahuila carbonate platforms,
de Graciansky, P.-C. & Vail, P. R. 1998. Chart 4: Early Cretaceous, northeastern Mexico. Geological
Mesozoic and Cenozoic sequence chronostratigraphic Society of America Bulletin, 111, 1010– 1029.
framework of European basins. In: de Graciansky, Lehmann, C., Osleger, D. A. & Montanez, I. P. 2000.
P.-C., Hardenbol, J., Jacquin, T. & Vail, P. R. Sequence stratigraphy of Lower Cretaceous
(eds) Mesozoic and Cenozoic Sequence Stratigraphy of (Barremian-Albian) carbonate platforms of northeast-
European Basins. Society for Sedimentary Geology, ern Mexico: regional and global correlations. Journal
Tulsa, OK, Special Publications, 60, 763–781. of Sedimentary Research, 70, 373– 391.
Hardie, L. A., Bosellini, A. & Goldhammer, R. K. Lehrmann, D. J. & Goldhammer, R. K. 1999. Secular
1986. Repeated subaerial exposure of subtidal carbon- variation in parasequence and facies stacking patterns
ate platforms, Triassic, northern Italy: evidence for of platform carbonates: a guide to application of
Downloaded from http://sp.lyellcollection.org/ by guest on July 5, 2013

340 L. A. HINNOV ET AL.

stacking-patterns analysis in strata of diverse ages and Schwarzacher, W. 2000. Repetitions and cycles in stra-
settings. In: Harris, P. M., Saller, A. H. & Simo, J. A. tigraphy. Earth-Science Reviews, 50, 51– 75.
(eds) Advances in Carbonate Sequence Stratigraphy; Scotese, C. R., Gahagan, L. M. & Larson, R. L. 1989.
Application to Reservoirs, Outcrops and Models. Plate tectonic reconstructions of the Cretaceous and
Society for Sedimentary Geology, Tulsa, OK, Special Cenozoic ocean basins. In: Scotese, C. R. & Sager,
Publications, 63, 187–225. W. W. (eds) Mesozoic and Cenozoic Plate Reconstruc-
Li, Y. X., Bralower, T. J. et al. 2008. Toward an orbital tions. Elsevier, Amsterdam, 27–48.
chronology for the early Aptian Oceanic Anoxic Event Snedden, J. W. & Liu, C. 2010. Compilation of
(OAE1a, 120 Ma). Earth and Planetary Science Phanerozoic sea-level change, coastal onlaps, and rec-
Letters, 271, 88–100. ommended sequence designations. AAPG Search and
Lowrie, W. 1990. Identification of ferromagnetic min- Discovery, article no. 40594, http://www.searchand
erals in a rock by coercivity and unblocking tempera- discovery.com/documents/2010/40594snedden/
ture properties. Geophysical Research Letters, 17, ndx_snedden.pdf
159– 162. Taner, T. H. 2003. Attributes revisited. Technical publi-
Malinverno, A., Erba, E. & Herbert, T. D. 2010. cation/ Rock Solid Images, Inc., Houston, TX,
Orbital tuning as an inverse problem: chronology of http://rocksolidimages.com/pdf/attrib_revisited.htm
the early Aptian oceanic anoxic event 1a (Selli Tauxe, L., Mullender, T. A. T. & Pick, T. 1996. Potbel-
Level) in the Cismon APTICORE. Paleoceanography, lies, wasp-waists, and superparamagnetism in mag-
25, PA2203, doi: 10.1029/2009PA001769. netic hysteresis. Journal of Geophysical Research,
Muxworthy, A. R. 1999. Low-temperature susceptibility 101, 571 –583.
and hysteresis of magnetite. Earth and Planetary Tauxe, L., Bertram, H. & Seberino, C. 2002. Physi-
Science Letters, 169, 51– 58. cal interpretation of hysteresis loops: micromagnetic
Ogg, J. & Hinnov, L. A. 2012. The Cretaceous Period. In: modeling of fine particle magnetite. Geochemistry,
Gradstein, F., Ogg, J., Ogg, G. & Smith, D. (eds) A Geophysics, Geosystems, 3, doi: 10.1029/2001GC
Geologic Time Scale 2012. Elsevier, Oxford, 793–853. 000280.
Paillard, D., Labeyrie, L. & Yiou, P. 1996. Macintosh Thomson, D. 1982. Spectrum estimation and harmonic
program performs time-series analysis. Eos Trans- analysis. Proceedings of the IEEE, 70, 1055–1096.
actions, American Geophysical Union, 77, 379. Tinker, S. W. 1982. Lithostratigraphy and biostratigra-
Pindell, J. L. 1993. Regional synopsis of Gulf of Mexico phy of the Aptian La Peña Formation, northeast
and Caribbean evolution. In: Pindell, J. L. & Mexico and southeast Texas, and the depositional
Perkins, B. F. (eds) Mesozoic and Early Cenozoic setting of the Aptian Pearsall-La Peña Formations,
Development of the Gulf of Mexico and Caribbean Texas subsurface and northeast Mexico: Why is there
Region, A Context for Hydrocarbon Exploration: not another Fairway Field? MSc thesis. University
Gulf Coast SEPM Foundation 13th Annual Research of Michigan, Ann Arbor, MI.
Conference Proceedings, 251–274. Wall, J. R., Murray, G. E. & Diaz, T. 1961. Geology
Pye, K. 1987. Aeolian Dust and Dust Deposits. Academic of the Monterrey area. Nuevo Leon, Mexico: Trans-
Press, London. actions– Gulf Coast Association of Geological
Röhl, U. & Ogg, J. G. 1998. Aptian– Albian Eustatic Sea- Societies, 11, 57– 71.
Levels. International Association of Sedimentologists, Weedon, G. P. 2003. Time-Series Analysis and Cyclos-
Wiley-Blackwell, London, Special Publications, 25, tratigraphy: Examining Stratigraphic Records of
95–136. Environmental Cyclicity. Cambridge University Press,
Schlager, W. 2005. Carbonate Sedimentology and Cambridge.
Sequence Stratigraphy. Society of Economic Paleon- Wilson, J. L. 1975. Carbonate Facies in Geologic Time.
tologists and Mineralogists, Tulsa, OK. Concepts in Springer, New York.
Sedimentology and Paleontology, 8, 200. Wilson, J. L. 1999. Controls on the wandering path
Schlager, W. 2010. Ordered heiarchy versus scale invar- of the Cupido Reef trend in northeastern Mexico.
iance in sequence stratigraphy. International Journal In: Bartolini, C., Wilson, J. L. & Lawton, T. F.
of Earth Sciences, 99(suppl 1), S139–S151. (eds) Mesozoic Sedimentary and Tectonic History
Schlanger, S. O. & Jenkyns, H. C. 1976. Cretaceous of North-Central Mexico. Geological Society of
oceanic anoxic events: causes and consequences. Geo- America, Boulder, CO, Special Papers, 340,
logie en Mijnbouw, 55, 179–184. 135–143.

You might also like