You are on page 1of 16

Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


journal homepage: www.elsevier.com/locate/palaeo

Record of Albian to early Cenomanian environmental perturbation in the


eastern sub-equatorial Pacific
J.P. Navarro-Ramirez a,b,⁎, S. Bodin a, U. Heimhofer c, A. Immenhauser a
a
Ruhr-Universität Bochum, Institut für Geologie, Mineralogie und Geophysik, D-44870 Bochum, Germany
b
Geological Survey of Peru (INGEMMET), San Borja-Lima, Peru
c
Leibniz Universität Hannover, Institut für Geologie, D-30167 Hannover, Germany

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

Article history: The present paper documents and discusses a new Albian–early Cenomanian carbon isotope (δ13Ccarb and
Received 12 July 2014 δ13Corg) curve from the subequatorial Eastern Pacific in Peru. Chemostratigraphic evidences for the expression
Received in revised form 22 January 2015 of the OAE1b set and for OAE1c and OAE1d are presented. This dataset is relevant inasmuch as previous work
Accepted 28 January 2015
is strongly biased towards study sites in North America (Western Interior Basin), in Europe (Tethys) and the Pa-
Available online 4 February 2015
cific realm. A comparison of the carbon isotope stratigraphy obtained in Peru with published sections from the
Keywords:
Central and Western Pacific, the Western Atlantic and Northern and Western Tethys reveals an overall good
Mid-Cretaceous agreement supporting the global nature of the isotope patterns described here. The δ13C from Peru record is
OAE constrained by biostratigraphic evidence and 87Sr/86Sr isotope stratigraphy using well-preserved oyster shells.
Chemostratigraphy Furthermore, we document the development of a heterozoan epeiric–neritic mixed carbonate–siliciclastic
Pacific ramp in the Western Platform of Peru and its corresponding sedimentary facies associations. This dataset was
Carbonates used to elucidate the complex interplay of climatic changes, nutrient supply, and platform drowning, leading
to the following conclusions: (i) an upper Aptian–lower Albian major change from siliciclastic-dominated to car-
bonate sedimentation coincided with the impact of the Kilian Level, (ii) a lower Albian incipient platform drown-
ing linked to the impact of the Paquier Level, (iii) A lower middle Albian major demise of neritic carbonate
production that coincides with the Leenhardt Level, followed by middle Albian condensed sedimentation that re-
ports prominent negative values in δ13Ccarb prior to the onset of OAE1c and (iv) finally, renewed carbonate ramp
production during the upper Albian–lower Cenomanian. The data shown here represent the foundation for fu-
ture work documenting the mid-Cretaceous of Peru and its implications for the palaeoceanography of the SE
subequatorial Pacific.
© 2015 Elsevier B.V. All rights reserved.

1. Introduction phases in marine life, changes in carbonate platform ecology and phases
of platform drowning (Hallock and Schlager, 1986; Wilmsen, 2000;
The Albian–early Cenomanian (mid-Cretaceous, ∼113–96.5 Ma; Ogg Föllmi, 2012; Krencker et al., 2014). Specifically, during the mid-Creta-
and Hinnov, 2012) witnessed a series of oceanic anoxic events (OAEs, ceous (Gale et al., 2011), OAEs took place (i) during the late Aptian–
Leckie et al., 2002; Heimhofer et al., 2004; Jarvis et al., 2006; Sprovieri early Albian (OAE1b, ∼ 114–109 Ma), (ii) during the early late Albian
et al., 2013; Lorenzen et al., 2013). One of the most prominent features (OAE1c, ∼ 102 Ma) and (iii) at the Albian–Cenomanian boundary
related to these events is represented by the accumulation of organic (OAE1d, ∼99.5 Ma).
matter in marine sediments (Schlanger and Jenkyns, 1976). These Much of the present understanding of mid-Cretaceous climate and
organic-rich sediments (black shales) can be traced in ocean basins its perturbations comes from the carbon-isotope proxy making use of
worldwide as individual, distinct horizons or as bundles with several bulk micrite (δ13Ccarb) and bulk organic matter (δ13Corg) sample sets.
layers enriched in organic carbon (Heimhofer et al., 2006; Emeis and Obtained chemostratigraphic sections from numerous basins world-
Weissert, 2009; Trabucho-Alexandre et al., 2012). This specific facies wide display excursions in the carbon isotope record that have signifi-
and related carbon isotope anomalies represent short-lived perturba- cance as time or event markers and shed light on processes involved
tions of the global carbon cycle (Jenkyns, 2010; Melinte-Dobrinescu (e.g., Erbacher et al., 1996; Bralower et al., 1999; Herrle et al., 2004;
and Bojar, 2010; Trabucho-Alexandre et al., 2010; Bodin et al., 2013). Jarvis et al., 2011; Krencker et al., 2014). Amongst these, OAE1b was a
OAEs are also associated with minor extinction and rapid turnover long-lived event lasting about 6.3 Myr, characterized by a bundle of
up to four black shale levels and associated perturbations in the carbon
⁎ Corresponding author. Tel.: +49 234 32 2325. cycle recorded as excursions in carbon isotopes (Herrle et al., 2003,
E-mail address: Juan.NavarroRamirez@rub.de (J.P. Navarro-Ramirez). 2004; Reichelt, 2005; Madhavaraju et al., 2013). In the Vocontian

http://dx.doi.org/10.1016/j.palaeo.2015.01.025
0031-0182/© 2015 Elsevier B.V. All rights reserved.
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 123

-70° -50° -50° -40° -30° -20° -10° 0° 10° 20° 30°
Basin (Western Tethys), these levels are represented by the uppermost
North Atlantic

x
Aptian Jacob Level and the lower Albian Kilian, Paquier and Leenhardt x
Ocean
10 N
x
levels. Here, the Kilian and Paquier levels are the most widespread, de- x
Palaeo- x x
posited during eustatic highstands (Bréhéret, 1994) and composed by Equ A
ator

x
-0 W
N D
marine organic matter with a low terrigenous input (Trabucho-

x
Alexandre et al., 2011). x
G O
-10
On the other hand, in the Tethys realm, the early late Albian OAE1c is ANDEAN

x
a black shale level characterized by abundant terrigenous organic mat- Farallon x BASIN
-20 Plate
ter (Amadeus Level; Coccioni and Galeotti, 1993; Galeotti, 1998; x
x x
Luciani et al., 2007). The chemostratigraphic pattern of this event is Pacific

x
-30 Ocean
not uniform, a feature most likely related to differences in sedimenta-

x
Phoenix

x
tion rates, the interaction of local and global processes and other, so Plate

x x
far not well constrained factors. In this sense, only a weakly developed -40
upper Aptian/lower Albian
negative δ13Ccarb shift has been reported in Central Europe for OAE1c x x Subduction zone

x
x
(Erbacher et al., 1996), whereas a prominent negative excursion in 1000 km x Area of study
-50
δ13Corg has been identified in Mexico (Bralower et al., 1999) or in sec-
tions on Japan (Takashima et al., 2010). At the Albian–Cenomanian tran- Fig. 1. Palaeogeographic map of Gondwana during the mid-Cretaceous (modified after
sition, OAE1d has been ascribed to a black shale level known as the Larson and Pitman, 1972; Torsvik et al., 2009; Moulin et al., 2010) indicating position of
Breistroffer Level in France (Gale et al., 1996). A long-lasting positive ex- what today is the Andean Basin. Black square denotes approximate position of study area.
cursion of δ13Ccarb within this interval has led several authors to suggest
globally significant organic-carbon burial (Nederbragt et al., 2001;
Schröder-Adams et al., 2012; Scott et al., 2013). It is likely, however, was accentuated overlying Lower Cretaceous siliciclastic units. Deposi-
that different OAEs have different driving mechanisms and differential tion was controlled by NNW–SSE-trending structures (Figs. 2 and 3).
organic matter-rich sedimentation in different localities during an The Paracas Massif, consisting of Precambrian continental basement,
event as reflected in different types of organic matter found in specific limits the western part of the Albian volcanic Huarmey Trough (Soler
black shale intervals (Kuypers et al., 2001). and Bonhomme, 1990). Further east, the Albian Huarmey Trough
Much of the present knowledge of the Albian–early Cenomanian evolved into an aborted marginal volcanic basin characterized by the
OAE records is biased towards data derived from sections in Europe deposition of basaltic pillow lavas of the Casma Group (Atherton and
(Tethys) and North America (Western Interior Basin). In contrast, very Webb, 1989). Volcanic activity in the Andean Basin was directly related
limited information on mid-Cretaceous OAEs is available from the east- to the break-up of the Gondwana arc, culminating with the opening of
ern sub-equatorial Pacific and generally, the western South American the South Atlantic Ocean (Torsvik et al., 2009; Moulin et al., 2010;
realm. A limited series of relatively well dated, albeit usually incomplete Winter et al., 2010; Fig. 1) and the emplacement of the Coastal Batholith
records of Albian to Cenomanian marine strata from the Pacific stems in the Huarmey Trough in the Late Cretaceous (Soler and Bonhomme,
from ocean drilling projects (e.g., Hess Rise, Shatsky Rise, Resolution 1990). In the middle part of the Andean Basin, the Western Platform
Guyot: Price, 2003; Robinson et al., 2004; Dumitrescu et al., 2006; was characterized as a back-arc basin, developed on an extensional tec-
Ando et al., 2008). Other outcrop-based studies from the Pacific realm tonic margin and activated during Jurassic–Cretaceous times (Jaillard,
have been reported from Japan (Nemoto and Hasegawa, 2011 and refer- 1987). Towards the southeast, the Western Platform was attached to
ences therein). Judging from available data, OAE1a and OAE2 seem to be the Marañon Massif, a generally submerged locally also emerged massif,
more represented in the Pacific region whilst other OAEs show a less separating the Western Platform and the Eastern Basin (Benavides-
pronounced record. Albian–early Cenomanian occurrences of organic- Caceres, 1956). The Eastern Basin was bound to the east by the
rich black shales seem to be less thick and rather patchy in distribution Brazilian shield and comprises deltaic coarse grained deposits (Figs. 2
(Robinson et al., 2004). In conclusion, the eastern sub-equatorial Pacific and 3).
is remarkably underrepresented and poorly constrained with respect to Given the scarcity of stratigraphic data on the Albian–early
continuous and well-dated chemostratigraphic reference sections for Cenomanian of Peru, a summary of the existing knowledge is given for
the mid-Cretaceous interval. reference (Fig. 4; Benavides-Caceres, 1956; Jaillard, 1987; Robert et al.,
In order to close this gap, a field-based project in northern Peru has 2009). In the northern Andes (Cajamarca region), the Lower Cretaceous
been undertaken and extended Albian–early Cenomanian sections are is represented by the Goyllarisquizga Group that encompasses the
here reported. The results shown indicate that OAE1b, 1c and 1d are re- Chimu, Santa, Carhuaz and Farrat formations (Benavides-Caceres,
corded in the Andean Basin of Peru. This paper has the following aims as 1956). The Goyllarisquizga Group is assigned to the Valanginites broggii
follows: to (i) document and interpret the sedimentology of well- Zone at the base and an Aptian age was assumed for the top (Benavides-
exposed Albian–early Cenomanian sections of the Andean Basin of Caceres, 1956). This group is overlain by Albian transgressive deposits
Peru; to (ii) provide a carbon-isotope reference curve from these sec- that resulted in shelf deposition of the Inca, Chulec, Pariatambo and
tions for the marginal SE Pacific and to (iii) discuss and correlate the Yumagual formations reaching the onset of the early Cenomanian.
Peruvian findings with coeval records documented from the Tethyan, The Inca Formation consists of iron-rich, sandy and marl–limestones
the proto-Atlantic and the Pacific domains. beds, assigned an early Albian age based on the ammonite
Neodeshayesites nicholsoni (Robert and Bulot, 2004; Robert et al.,
2. Regional tectonic and stratigraphic setting 2009). The Inca is unconformably overlain by the Chulec Formation
with a discontinuity surface separating the two units. The Chulec For-
During the mid-Cretaceous, the Andean Basin was located within mation is characterized by marl–limestone alternations with a very
the sub-equatorial humid belt of the Southern Hemisphere (Hay and abundant and diversified outer shelf fauna (Jaillard, 1987). Numerous
Floegel, 2012) and separated from the Pacific Ocean by a volcanic arc ex- ammonites have been reported and were assigned to the Knemiceras
tending some 10,000 km from South America to the North Scotia Ridge raimondii Zone (Robert and Bulot, 2004; Robert et al., 2009), indicating
in the South Atlantic (Fig. 1; Larson and Pitman, 1972; Atherton and a middle early Albian–early middle Albian age. The Chulec Formation is
Aguirre, 1992). Owing to a long-term transgression during the Albian conformably overlain by the Pariatambo Formation.
(Pindell and Tabbutt, 1995; Robert, 2002), broad portions of the Andean The Pariatambo Formation is characterized by fossiliferous, black, bi-
Basin were flooded and mixed carbonate–siliciclastic sedimentation tuminous, fetid marly limestones facies and includes fine lamination
124 J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

80°W 75°W 70°W

v
N COLOMBIA
v ECUADOR
v

v
v v
5°S v
v
v
v
BRAZIL

We
v
v

st
A v

er
Ea
Pa

n P
v
10°S st
rac
v
x er
n B

lat
x
v a
as

si n

BO
Ma
for
PACIFIC v
ra
x

LIV
LIMA
Ma

ño x x
m
OCEAN

IA
v
n M x x
ss

v a ss
i
if

v
v v x

f
0 100 200 km x
u
H

x x
deltaic sandstone ar
15°S x m x
inner-shallow ramp x e
y

middle-outer ramp Tro


outer deep ramp ug
v Albian volcanic arcs v
x Palaeozoic-Precambrian v v
basements v

Fig. 2. Mid-Cretaceous palaeogeographic map of Peru (based on Pindell and Tabbutt, 1995; Robert and Bulot, 2004). Black square indicates study area. Note size of study area relative to the
vast dimensions of the Western Platform. A–A′ denotes position of transect in Fig. 3.

and siliceous intervals. Ammonites and planktonic foraminifera bioclastic grey marl–limestones of the Yumagual Formation containing
are common, evidencing the Oxytropidoceras carbonacrium and few ammonites. Bivalve biostratigraphy places the Yumagual Formation
Prolyelliceras ulrichi Zones assigned to a middle Albian age (Robert within the Ostrea scyphax and Exogyra mermeli Zones, indicating a late
et al., 2009). The Pariatambo Formation grades upsection into nodular middle Albian–early Cenomanian age (Benavides-Caceres, 1956).

A Andean Basin A´
gh ey

si on

ld an
fo rn
si as

in rn
an c

ou rm

rm

ie ili
as añ
at te
as c
ce ifi

as te
M ara

sh raz
f

f
Pl es
Tr ua
O ac

M ar

B as
W

M
H

B
P

Albian volcanic arcs outer ramp middle ramp inner ramp deltaic sandstones

Distal Proximal
Study area
Sea level
UPPER CEOUS
CRETA
L
UA
Y U MAGMBO
C
G asma PARIATA
C
r o up
low CHULE
er K 0 100 200 km

limestone pillow lavas


marl, locally sandstone or claystone terrestrial or marine volcanics
deltaic sandstone intrusions
coarse grained continental deposits faults

Fig. 3. Hypothetical transect across the Andean Basin during Neocomian–Coniacian times (based on Benavides-Caceres, 1956; Atherton and Webb, 1989; Jaillard, 1987) with inferred
Western Platform ramp morphology. Refer to approximate position of transect in Fig. 2 (A–A′).
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 125

87 86
Age Sr / Sr
Stage Formation Main fossil reported Zones
(Ma) 0.70725 0.70735 0.70745

Cenomanian
early
Mariella lewesiensis.,
Nummoloculina heimi Exogyra mermeli

100

Ostrea scyphax Hua 199


Yumagual Engonoceras sp.,
MIDDLE CRETACEOUS

Paraturrilites lewesiensis.,
Ostrea scyphax
late Sharpeiceras occidentale

105
Albian

Oxytropidoceras carbonarium,
Venezoliceras sp.,
Branceras sp., Lyelliceras lyelli, Oxytropidoceras
middle

Pariatambo Dipoloceras sp.,Hedbergellae carbonacrium


Puy 91.3
110 Lyelliceras lyelli, Knemiceras raimundi,
Douvilleiceras monile, Protanisoceras Knemiceras
Chulec
blancheti, Parengonoceras sp., raimondii
early

Beudanticeras sp.,Hamites sp.,


Platiknemiceras sp.,Douvilleiceras rex, N. nicholsoni,
Inca
Ap

Neodeshayesites umbilicostatus umbilicostatus


l

L o w e r C r e t a c e o u s G o y l l a r i s q u i z g a G r o u p

Fig. 4. Cretaceous chronostratigraphy of the Northern Andes. Note Goyllarisquizga Group, Inca, Chulec, Pariatambo and Yumagual formations and corresponding biostratigraphic frame-
work (Benavides-Caceres, 1956; Robert and Bulot, 2005; Robert et al., 2009). Variations in seawater 87Sr/86Sr values during early Albian–early Cenomanian times are shown (after Howarth
and McArthur, 1997; McArthur et al., 2001). Timescale is from Ogg and Hinnov (2012). Horizontal lines indicate strontium isotope values of best-preserved oysters used in this study.
Numerical ages are derived following the procedures of Howarth and McArthur (1997).

3. Methods and materials Cenomanian age. Sections were located along the road from Cajamarca
to Celendin (Fig. 5). In total, ca. 450 m of section has been logged. These
3.1. Field work and thin-section microscopy field data, obtained rock samples and thin-sections provide the funda-
ment for the petrographic and facies interpretations presented here. In
Two well-exposed sections (Pulluicana and Huameripashga) were analogy with the nomenclature and approach presented in
chosen for their well-exposed strata of the early Albian–early Immenhauser et al. (2004), non-skeletal and skeletal components

756000 768000 780000 792000 804000 816000 828000


Mara

9252000
ñon river
PE
RU

9240000 Lima CELENDIN

N
9228000

HUAMERIPASHGA
9216000

CAJAMARCA PULLUICANA
9204000

9192000

0 12 24 km

Quaternary deposits Lower Cretaceous siliciclastic deposits Fault


Tertiary volcaniclastic deposits Triassic–Jurassic marine deposits Road
Upper Cretaceous marine deposits Precambrian continental basement Section
Mid–Cretaceous marine deposits Subvolcanic rocks measured

Fig. 5. Geological map of the Cajamarca region (northern Andes) showing main structural elements and location of Pulluicana and Huameripashga study areas.
Modified after INGEMMET (2006).
126 J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

were analysed semi-quantitatively with numbers indicating their rela- distilled water via centrifuge and dried at 40 °C. The δ13Corg analysis
tive abundance: 0 = absent, 1 = present, 2 = frequent, 3 = abundant, was conducted using a Finnigan MAT 252 at the stable isotope laborato-
4 = dominant. Oyster shells were collected as potential archive material ry of the Leibniz University Hannover, Germany. All isotope results are
for radiogenic strontium-isotope stratigraphy. Limestone classification reported in per mil (‰) relative to the Vienna-Pee Dee Formation bel-
is according to Dunham (1962). Discontinuity surfaces, mainly marine emnite (V-PDB) standard in the conventional manner.
hardgrounds, were recorded stratigraphically and traced laterally be-
tween the two sections. Combined with under- and overlying facies, 3.4. Strontium-isotope analysis
discontinuities acted as pining points for a first attempt to understand
the sequence stratigraphic evolution of the study area. Emphasis must Sample powders of low-Mg calcite shells (n = 5) of oysters were
be laid on the fact that sequence stratigraphy as based on a limited num- taken by use of tungsten drill bits. The required sample quantity of
ber of sections is at best regional in nature but serves as an anchor point 500 ng was determined as based on the strontium content of each sam-
for future work. ple. The Sr fraction was separated twice using quartz glass columns
filled with BioRad ion exchange resin AG50W-X8 according to column
3.2. Assessing the diagenetic alteration of shell material calibration pattern. The 87Sr/86Sr ratio was determined with a 7 collec-
tor Thermal Ionisation Mass spectrometer (TIMS) MAT 262 at the
3.2.1. Cathodoluminescence microscopy Ruhr-University Bochum, Germany, in a 3 collector dynamic mode. As
Cathodoluminescence (CL) microscopy examination of selected standard reference materials NIST NBS 987 and USGS EN–1 were cho-
outer shell layers of five oyster specimens were investigated for their sen. The total blank for Sr isotope analysis, including chemical separa-
CL pattern in order to avoid diagenetically altered shell material. This tion and loading blank is 1.5 ng. The repeatability was tested with the
was carried out with the ‘hot cathode’ CL microscope (type HC1-LM) USGS EN–1 modern bivalve carbonate, which passed through the
at the Ruhr-University Bochum, Germany. The acceleration voltage of same procedures as all samples. The average 87Sr/86Sr value is
the electron beam is 14 kV and the beam current is set to a level gaining 0.709160 ± 0.000027 2σ (n = 209). The reproducibility of Sr measure-
a current density of ~9 μA mm−2 on the sample surface. Refer to Christ ments represented by NBS987, which was directly loaded onto a Re fil-
et al. (2012) for details on the analytical procedure. ament is 0.710240 ± 0.00034 2σ (n = 233). Refer to Huck et al. (2011)
for more details of the analytical procedure.
3.2.2. Trace and major element analysis
Oyster shells that passed initial cathodoluminescence inspection 4. Data description and interpretation
were subsequently analysed for major and trace elemental composition
at the Ruhr-University Bochum, Germany. Aliquots of the powder sam- 4.1. Facies associations and depositional environments
ples (1.35–1.65 mg) were analysed by inductively coupled plasma
atomic emission spectroscopy (ICP-AES) for strontium, magnesium, The mid-Cretaceous facies belts along the Andean Basin of South
iron and manganese (Table 1). Refer to Huck et al. (2011) for details America crop out over a north–south extension of about 2000 km
on the analytical procedure. (Figs. 1 and 2). Hence, despite the wide regional extent of sections mea-
sured in the context of this study, the observational window represent-
3.3. Carbon isotope stratigraphy ed by our data is dwarfed by the size of the Cretaceous Western
Platform. Having said this, the direct match of coeval portions of mea-
3.3.1. Bulk micrite (δ13CCarb) sured chemostratigraphic sections argues for an over-regional signifi-
Carbonate bulk rock specimens were sampled by use of tungsten cance of the data shown here. More fieldwork in neighbouring areas
drill bits. Whilst drilling the carbonate powder, areas rich in sparry ce- will support this working hypothesis or claim revision.
ment, large bioclasts and diagenetic calcite vein material were avoided. The data obtained so far and previous work point to a mixed, carbon-
In order to reveal the variability of carbon-isotope composition within a ate–siliciclastic ramp with decreasing argillaceous influence towards
single rock sample, several sub-samples were drilled from 20% of all the the west (i.e., Pacific-wards) pinching out against the Albian volcanic
hand specimens collected. Refer to Immenhauser et al. (2005) for more arc (Figs. 1 and 2). The most proximal settings represented by coastal,
details of the analytical procedure. Carbon-isotope analysis of 123 deltaic and tidal flat facies in the eastern part of the platform
micrite samples was performed using a Thermo Finnigan MAT delta-S (present-day eastern Peru and western Brazil, Fig. 2) were not visited
mass spectrometer at the isotope laboratory of the Ruhr-University in the context of this project. Integrating previous work (e.g.,
Bochum, Germany (see Appendix A). Repeated analyses of certified car- Benavides-Caceres, 1956; Robert et al., 2009), three main depositional
bonate standards (NBS 19, IAEA CO–1 and CO–8) and internal standards environments, each with a number of standard facies types, are here
show an external reproducibility of ≤ 0.06‰ for δ13Ccarb. The δ13Ccarb established for the subtidal to outer ramp (Fig. 6 and Table 2). The doc-
values are expressed on a per mil (‰) basis relative to the Vienna-Pee umentation of these fundamental sedimentological data is important
Dee Belemnite standard (V-PDB). due to the lack of accessible previous work. Estimates of the approxi-
mate bathymetry of corresponding depositional environments are
3.3.2. Bulk organic matter (δ13Corg) based on previous works dealing with comparable depositional envi-
Carbon-isotope analyses were performed on 84 bulk organic matter ronments (Yanin, 1983; Lukasik et al., 2000; Immenhauser, 2005, 2009).
samples (see Appendix A). Two grammes of sample powder was placed
in 50 mL centrifuge tubes and 6N HCl was added. This procedure was re- 4.1.1. Shallow subtidal inner ramp setting
peated until no carbonate reaction was visible. The samples were centri- The shallow subtidal inner ramp depositional environment com-
fuged and the supernatant removed, the residues were rinsed with prises three facies types: 1a, 1b and 1c (Fig. 6 and Table 2). The most

Table 1
Selected elemental and 87Sr/86Sr data of low-Mg calcite oysters from the Cajamarca region (Western Platform, Peru) and derived numerical ages.
87
Locality Sample no. Sr Mg Fe Mn Sr/86Sr 2σ mean Age [Ma]

[ppm] (×10−6) Min Mean Max

Huameripashga Hua 199 380 2598 720 76.6 0.707435 7 97.93 101 104.12
Pulluicana Puy 91.3 825 873 147 6.3 0.707380 6 108.87 109.36 109.74
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 127

Outer ramp Middle ramp: Inner ramp zone:


(~30–50 m) (10–30 m) (<5 m)

below storm wave-base below fair-weather base shallow subtidal zone

mud/wackestones rud/float/grainstones marl/mud/wacke/packstones

redox condictions open marine shallow waters


Sediment input
Storms

Sea level

and reworking
erosion 0 100km
oyster bioherms
3 2 2 1 1
(3a) (2c) (2a–2b) (1c) (1a–1b)
Finely-bedded, Thinly- Well-preserved oysters. Imbricated, chaotic, Bioturbated, Scarce fauna, faint
bituminous, bedded, Commonly bioturbated. and fragmented bioclastic banks, bioturbation, erosional
dark-grey, sulphides, marly, skeletal elements. diverse fauna, surfaces.
Chondrites. bioturba- Parallel lamination, erosional surfaces.
ted. intraclasts.

Fig. 6. Epeiric ramp model for the Western Platform in Peru exposed to occasional storm events and depositional environments recorded in the Inca, Chulec, Pariatambo and Yumagual
formations of the Cajamarca region. Facies types (1a through 3a) are indicated. Refer to Table 2 for more details. FWWB = fair-weather wave base; SWB = storm-weather wave base.

proximal deposits are composed of grey argillaceous mudstone and oc- amounts (Facies 1b; Fig. 8A). Evidence for bioturbation is limited and
casionally iron-rich sandstone, representing intervals of two to five me- primary sedimentary structures such as plane bedding are well pre-
tres thick (upper portions of Inca Formation). Scarce fauna occurs as served (Fig. 7A).
undifferentiated shell debris and flora as plant remains (Facies 1a). In Towards more open marine settings, Facies type 1b grades into
more distal sub-environments, the beds show thickening-upward pack- thicker beds, characterized by grainstone and occasionally floatstone fa-
ages and the facies is characterized by packstones and argillaceous cies and reaching 1 to 3 m in thickness (Facies 1c, lower portions of
wackestones, exhibiting a nodular fabric due to an increased argilla- Chulec Formation). Facies 1c is typically rich in mixed and fragmented
ceous content, intraclasts and oncoids are also observed in variable shell debris, mainly of oysters and gastropods and echinoderms

Table 2
Overview of facies classification and interpretation. Numbers indicate the relative abundance of non-skeletal and skeletal components: 0 = absent, 1 = present, 2 = frequent, 3 =
abundant, 4 = dominant.

Facies Colour Facies/key Texture Skeletal and non–skeletal Depositional Facies associations and
code code biota components environment depositional environment

Restricted fauna m Plant remains 1–2, quartz 2–3


1a Shallow subtidal
with quartz
1

Argillaceous m, M–W Micrite envelopes 1–2, peloids 1–2, oyster


1b Shallow subtidal Shallow subtidal
restricted fauna 1, oncoids 1–2, shell debris 1
inner ramp
Peloids 1, oyster 2–3, gastropods 1–2, shell
Argillaceous W, P–G, debris 1–2, oncoids 1, echinoderms 1–2,
1c Shallow subtidal
diverse fauna F benthic foraminifera 1–2

Peloids 1, oyster 1–2, gastropods 1–2,


Diverse fauna– P–G, F shell debris 2–3, echinoderms 2–3,
2a Open marine ramp
high energy benthic foraminifera 1–2
Open marine–
2 middle ramp
Oyster 2–4, shell debris 2–3,
2b Oyster floatstone F–G, R Open marine ramp
echinoderms 1–2

Diverse fauna– Shell debris 1, echinoderms 1–2,


2c M–W, P Open marine ramp
low energy planktonic foraminifera 1–2

Dark grey, Shell debris 1, echinoderms 1,


3a M Outer ramp 3 Outer ramp
muddy facies planktonic foraminifera 1
128 J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

Fig. 7. A): Outcrop image of thickly-bedded Yumagual Formation. B) Bioclastic rudstone of the Yumagual Formation, characterized by oysters in life position in shell debris host facies
(Facies 2a). C) Oyster specimen used for strontium analyses in this study (Facies 2b). D) Discontinuity surface (SB1) that marks the contact between medium-scale sequences (MsS)
1 and 2. Note olistolithic block. MsS 2 displays thin dark-grey to thick-bedded limestones of the Pariatambo Formation.

(Fig. 8B). Bioturbation is locally more intense and the bedding blurred. storm events (Facies 1c). In the absence of corals, calcareous green algae
Abundant minor discontinuity surfaces are observed. These are charac- or ooidal facies, an estimate of the palaeo-bathymetry of the distal inner
terized by FeO mineralization and increased levels of bioturbation of un- ramp is difficult. In comparison to data shown in Yanin (1983) water
derlying rocks. depths from near-sea level to about 5 m are tentatively estimated for
The lithological and palaeontological characteristics of the inner- deposits of facies type 1c (Fig. 8B). In this sense, it seems likely that fa-
most subtidal ramp facies exposed in the study area supports a restrict- cies types 1a through c are not differentiated by bathymetry but rather
ed, low-energy environment, where punctuated tempestite intervals, through their decreasing level of clastic influx and increasingly marine
abundant terrigenous influx including plan remains and the scarcity of water masses.
a marine ichnofacies indicate a significant level of continentality. Esti-
mates of the palaeo-bathymetry as based on comparable settings else- 4.1.2. Open marine middle ramp setting
where point to water depths of a few metres only (Facies 1a). The more open marine, middle ramp depositional environment
Limestones deposited in the more distal inner ramp setting, i.e., to- comprises facies types 2a, 2b and 2c (Fig. 6 and Table 2). Deposits of
wards the west, display a more diverse fauna and much decreased ter- the mid-ramp setting are characterized by packages of oyster bioherms,
rigenous influx suggesting a protected, shallow subtidal environment interbedded by rare grainstone units yielding various skeletal elements
under low hydrodynamic energy near the expectedly very shallow and intraclasts (Facies 2a; Fig. 7B). Oysters are a dominant faunal ele-
fair-weather wave base (Immenhauser, 2009; Facies 1b). The presence ment and associated with echinoderms and occasional ammonites.
of transported, fragmented and worn skeletal grains points to periodical Skeletal elements are locally imbricated, elsewhere bioclasts are
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 129

Fig. 8. A): Packstone of Chulec Formation composed of oncoids and variegated shell debris (Facies 1b). B) Bioclastic floatstone of the Chulec Formation, showing mixed and fragmented
skeletal debris (Facies 1c). C) Wacke- to packstone of the Chulec Formation, showing some gastropods, echinoderms and unspecified shell debris (Facie 2c). D) Mudstones with Chondrites
burrows of the Pariatambo Formation (Facies 3a).

arranged in lenses and display a chaotic fabric of non-sorted and reach of storm waves in relatively dys- to anoxic water conditions at
fragmented bioclasts. Oyster packstones represent bioherms with di- the sea floor.
mensions of between 5 and some tens of metres and thicknesses of up In modern, open storm exposed oceanic margins, the effective storm
to one metre (Facies 2b; Fig. 7C). Basin-wards, beds display thinning wave base is in the order of 150 m or less (Immenhauser, 2009). Con-
and are represented by bioturbated mud- to wackestones, less com- sidering the epeiric shelf and the potential barrier effects of the Albian
monly by packstones. Faunal elements consist of gastropods, echino- volcanic arc to the west (Atherton and Webb, 1989), however, this
derms and planktonic foraminifera and of unspecified shell debris depth range seems unlikely due to wave base frictional energy loss
(Facies 2c; Fig. 8C). and a depth range of 30–50 m is suggested in a tentative manner. This
Generally, oyster-debris facies witness storm hydrodynamics that is the typical depth range of smaller, epeiric–neritic basins that might
eroded, entrained and re-deposited oyster bioherms that were other- represent some sort of an analogue of the broad Peruvian shelf in the
wise situated beneath the effective fair-weather wave base (Facies mid-Cretaceous but the modern world provides no suitable analogues
2a), i.e., the depth at which wave-orbitals are still capable of moving to the Cretaceous Western Platform of Peru.
sediment particles (Immenhauser, 2009). According to Yanin (1983),
oyster bioherms in mid-ramp settings are often situated at palaeo- 4.2. Sequence stratigraphic interpretation
water depths of between 10 and 20 m (Facies 2b). In modern, wave-
exposed oceanic margins the average effective fair-weather wave base Acknowledging the fact that the present data set covers a compara-
is in the order of 30 ± 15 m. Given the broad epeiric ramp setting stud- ble small portion of the vast Western Platform of the Cretaceous of Peru,
ied here, fair-weather and swell waves from the Cretaceous Pacific a first-order assessment of the sequence stratigraphy of these sections is
Ocean probably lost much of their energy due to wave base-seafloor in- here presented. The significance of these data lies fundamentally in the
teraction and the shallower end-member of this bathymetric range is fact that they represent the foundation for future work from this poorly
expected (Keulegan and Krumbein, 1949). Moreover, the significance documented part of the world and allow for a comparison with se-
of the Albian volcanic arc to the west is difficult to quantify in terms of quence stratigraphic interpretations from neighbouring basins. The fa-
wave climates. The combination of oyster bioherms, combined with cies analyses of the Pulluicana–Huameripashaga composite section
gastropods and echinoids, the presence of planktonic foraminifera and indicates changes in accommodation space of a least two different or-
ammonites may indicate an open marine setting with a water depth be- ders of magnitude. Essentially, the Inca, Chulec, Pariatambo and
tween ∼10–30 m. Yumagual formations represent a large-scale sequence (LsS1) with a
minimum stratigraphic thickness of 450 m (Figs. 9 and 10). As the strat-
4.1.3. Outer ramp setting igraphic top of the Yumagual Formation is not reached in the outcrops
The outer ramp setting is characterized by facies type 3a (Fig. 6 and visited, LsS1 is perhaps stratigraphically thicker than documented
Table 2) and is located below the reach of the effective storm wave base. here. Overall, large-scale sequence 1 is made up by three medium-
The facies obtained primarily consists of mudstones in dm-thick beds scale sequences with thicknesses ranging between 125 and 150 m, com-
and yields traces of sulphide deposits (pyrite content). The carbonate fa- monly delimited by well-marked discontinuity surfaces indicating
cies is dark grey in weathering colour (Fig. 7D). Chondrites-type biotur- changes in the depositional system including abrupt shifts of facies
bation is common (Fig. 8D). The faunal composition includes planktonic belts (Fig. 9).
foraminifera and echinoderms (Facies 3a). Fine-grained deposits char- Medium-scale sequence 1 (MsS1) overlies the argillaceous mud-
acterized by dark grey colouring, fine lamination, high pyrite content, stone and iron-rich sandstone alternations of the Inca Formation. At
chondrites and scarce fauna may indicate an environment below the the base, the transgressive system tract is represented by an alternation
130
Formation
Substage
Pulluicana (lower) - Huameripashga 13 13
C org (‰ V-PDB) Carbonate
C carb(‰ V-PDB) Sea-level

Stage

Exogyra mermeli Zone


(upper) composite section -3 -2 -1 0 1 2 -28 -27 -26 -25 50m 30 10 0 Events factory

Medium-scale Sequence 3 (MsS3)


250m deeper
CENOMANIAN
early
225 Pe9b

200
Hua 199
87
Sr/ 86 Sr= 0.707435± 7x10-6
Pe9 Breistoffer OAE1d?
MFI 175

J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137


YUMAGUAL
Ostrea scyphax

150
late

SB2 Pe9a
125
Medium-scale Sequence 2 (MsS2)

100
Large-scale Sequence 1 (LsS1)

75

50 Pe8
Oxytropidoceras carbonacrium
ALBIAN

MFI Pe8
Pe7 Pe7
25
PARIATAMBO

175m

SB1 Leenhardt
middle

0 Pe6
Pe6
Medium-scale Sequence 1 (MsS 1)

mMWPGFR 150

MFI 125

Pe5 Pe5 OAE1b


87
Puy 91.3 Sr/ 86Sr= 0.707380± 6x10-6 set
-6
CHULEC

75
N. nicholsoni K. raimondii

Pe4
50 Pe4
Pe3 Paquier
Pe3
early

25
Pe2
INCA

X Pe1
0 Kilian
0 1 2 -28 -27 -26 -25 50m 30 10 0
mMWPGFR

Fig. 9. Regional sequence stratigraphic interpretation, carbon-isotope stratigraphy and relative sea-level record of the Pulluicana–Huameripashga composite section with indication of Albian OAEs and corresponding sub-levels. Pe1–Pe9 refers to
chemostratigraphic segments defined for the Peruvian reference curve. Data points shaded grey correspond to OAEs based on bio- and chemostratigraphic data. Note position of 87Sr/86Sr data. Biostratigraphic data by Benavides-Caceres (1956)
and Robert et al. (2009).
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 131

chondrites m = marl
Restricted fauna with quartz Depositional
bioclasts (undiff.) M = mudstone Sequences
Argillaceous restricted fauna W = wackestone
bivalves (undiff.)
Argillaceous diverse fauna oysters P = packstone
HD

Medium-scale
G = grainstone

Large-scale
Diverse fauna-high energy gastropods
planktonic forams F = floatstone
Oyster floatstone R = rudstone MFI
echinoderms
Diverse fauna-low energy nodular aspect
silicification TD
Thinly-bedded turbidites
oncoids
Dark grey, muddy facies
peloids HD = highstand deposit
Siliciclastics intraclasts MFI = maximun flooding interval
Heterozoan assemblages (calcite) Sr-samples TD = transgressive deposit

Incipient demise or condensation - undifferentiated bioturbation


discontinuity surface

Fig. 10. Legend for Fig. 9, denoting colour codes for different facies types and corresponding carbonate factory.

of wacke- and mudstone beds intercalated with some floatstone– Medium-scale sequence 2 (MsS2), i.e., the transgressive systems
packstone units (Chulec Formation). It is topped by a discontinuity sur- tract comprises wackestones, locally with pack- float- and rudstone al-
face bearing field evidence of a marine hardground, which marks the ternations and some rare calciturbidites. The maximum flooding inter-
maximum flooding surface and, thus, the contact between the Chulec val is probably represented by thinly-bedded, black argillaceous facies
and the Pariatambo formations. The hardground is overlain by more cal- and, thus, the contact between the Pariatambo and the Yumagual for-
careous and thicker bedded highstand deposits comprising wacke- to mations. Upsection, this facies grades into highstand deposits character-
floatstone units that are often bioturbated. A regionally significant dis- ized by packstone with rud- to floatstone alternations (lower to middle
continuity (SB1) characterized by the impregnation with secondary portions of the Yumagual Formation). At the top, discontinuity surface
iron-oxides and intense bioturbation is recognized in both, the SB2 is characterized by impregnation and intense bioturbation and
Huameripashga and Pulluicana sections respectively. On this disconti- marks the top of MsS2 and represents the transgressive surface at the
nuity surface an olistolithic block is observed (Fig. 7D) that may indicate base of MsS3 (Fig. 9).
local tectonic faulting along the northern Andes (Jaillard, 1987). Medium-scale sequence 3 (MsS3) comprises the upper exposed
In terms of its sequence stratigraphic significance, SB1 marks the portions of the Yumagual Formation. Here, transgressive deposits are
transgressive surface of the MsS2 (Fig. 9). Due to the lack of evidence represented by alternations of packstones and argillaceous nodular
for prolonged subaerial exposure of SB1, such as karstification, wackestones, overlain in turn by a minor marine discontinuity that
bleaching or saw-tooth-shaped carbon and oxygen-isotope excursions, may well define the maximum flooding surface. Highstand deposits
the medium-scale sequences shown in Fig. 9 might represent are represented by thickly-bedded, grey wackestones that grade
parasequences topped by marine transgressive surfaces. upsection into even thicker bedded, bioturbated grain- and floatstones

A B

500 µm 500 µm

C D

500 µm 500 µm

Fig. 11. A) Cathodoluminescence photomicrograph of oyster shell, dark blue luminescence indicates well-preserved shell material. B) Same image under crossed polarizers.
C) Cathodoluminescence photomicrograph of altered shell. D) Same image under crossed polarizers.
132 J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

facies. The top of the Yumagual Formation is not exposed in the study The δ13Ccarb section starts with an abrupt negative shift reaching
area. values as low as + 0.1‰ (Pe3), continued by a recovery phase (Pe4).
Thereafter, Pe5 shows a plateau that oscillates between 0 and +2.3‰,
5. Chemostratigraphy Pe6 presents a sharp δ13Ccarb decrease reaching −0.1‰, followed by rel-
atively invariant values (Pe7). Pe8 displays a pronounced negative shift
5.1. Radiogenic strontium isotope stratigraphy in δ13Ccarb ratios, reaching a minimum of −3‰ (the lowest values mea-
sured in these sections). Upsection, δ13Ccarb values display a gradual re-
In light of limitations related to available biostratigraphic data (e.g., turn to pre-excursion values (Pe9a), followed by a plateau with values
Benavides-Caceres, 1956; Robert et al., 2009 and references therein) in the order of between +1.3 and +2.1‰ (Pe9b).
from the Inca, Chulec, Pariatambo and Yumagual formations, it is
critically important to apply independent stratigraphic tools in order 6. Discussion
to better constrain the chronological framework (Fig. 4 and Table 1).
Strontium-isotope (87Sr/86Sr) analyses of well-preserved low-Mg 6.1. Large-scale depositional setting of the Western Platform
calcite biominerals (rudists, belemnites, brachiopods and oysters) rep-
resent a powerful tool in chemostratigraphy (Huck et al., 2011; Controls that play a main role in the type of mid-Cretaceous neritic
Mutterlose et al., 2014) that is well adapted for the mid-Cretaceous. carbonate deposition (Lukasik et al., 2000; Schlager, 2005) include sea-
The first and most important step, however, is the evaluation of well- water temperature, degree of restriction, nutrient levels or latitudinal
preserved biogenic carbonate material for further analysis. diversity gradients. Data shown here document the predominance of
The outer shell layer of a series of macro- and mesoscopically well- heterozoan skeletal elements in the Western Platform of Peru. In tropi-
preserved oysters has been investigated for their diagenetic overprint cal settings, heterozoan carbonate secreting organisms are perhaps
(Fig. 11). Intact valves and/or fragments of oysters display various levels most common in open marine and relatively cool seawater settings
of preservation. Cathodoluminescence microscopy revealed well- (Philip and Gari, 2005), or where enhanced nutrient levels prevail
preserved valves with formerly organic-rich layers now replaced by (Halfar et al., 2004). Hence, a potential cause for the dominance of a
secondary luminescent phases (Fig. 11A and B), whilst others display heterozoan association in the palaeo-tropical setting of Peru is best
growth increments replaced by secondary luminescent carbonate found in the lower to middle Albian transgression that likely favoured
phases (Fig. 11C and D). The latter specimens were not further an ingression of nutrient-rich oceanic water masses from the Pacific or
considered here. Well-preserved valves were selected for further geo- the Atlantic. Nevertheless, evidence for the establishment of an open
chemical analyses (Sr, Mg, Fe and Mn elemental abundances). High marine circulation in the Andean basin is indicated by the occurrence
manganese and iron concentrations (N 100 ppm) in calcitic oyster of ammonite taxa of Tethyan affinities (Robert and Bulot, 2005). Accord-
shell are used to determine recrystallization under reducing conditions ing to Trabucho-Alexandre et al. (2010), nutrient-rich surface currents
(Huck et al., 2011 and references therein). Two out of five selected oys- were flowing along the southern margin of the proto-Atlantic into the
ter specimens fulfilled these boundary conditions and were used for Pacific via the northern margin of South America. It is at least conceiv-
chemostratigraphy. As shown in Table 1, these shells yield Sr elemental able that some of these water masses were flooded onto the Andean
abundances ranging from 380 to 825 ppm and Mn elemental abun- epeiric ramp. Moreover, the Palaeo-Pacific Ocean to the east represents
dances ranging from 6 to 76 ppm. Iron abundances exhibits more than a potential upwelling zone for cooler, nutrient-rich bottom waters sim-
100 ppm, an observation that is in agreement with the mild degree of ilar to the present-day situation. However, the Cretaceous ocean water
alteration observed under the cathodoluminescence microscope being structure and current patterns were probably considerably different
spatially limited to formerly organic-rich shell portions. from those of the modern Pacific (D'Hondt and Arthur, 1996; Hay and
Numerical ages are derived from mean values of theses best- Floegel, 2012) and differences in the plate tectonic setting were of im-
preserved samples using the ‘look up’ table of McArthur et al. (2001). portance, too. Finally, both the Marañon Massif to the east, as well as ex-
The upper and lower age limits were obtained by adding two standard posed continental basement in present-day Brazil in the east (Figs. 2
errors of the mean values of isotopic results to the statistical uncertainty and 3) formed an important and spatially large source area for continen-
of the seawater curve (Howarth and McArthur, 1997). The results indi- tal runoff delivering freshwater, nutrients and terrigenous influx. Judg-
cate numerical ages of 109 ± 0.4 Ma (early middle Albian) for the mid- ing from the argillaceous facies in the innermost exposed ramp settings
dle part of the Chulec Formation and 101 ± 3 Ma (late Albian–early (facies type 1a; Fig. 6 and Table 2), Albian transgressive pulses might
Cenomanian) for the Yumagual Formation (Huameripashga section; have rework clay and silt facies in the flooded coastal settings.
Table 1). Combined with the results of the detailed carbon-isotope Taking these potential nutrient-sources into account, the overall sce-
curve documented below and previously published biostratigraphic nario of a mixed carbonate-siliciclastic ramp – protected perhaps locally
data (e.g., Benavides-Caceres, 1956; Robert et al., 2009 and references by the Paracas structural high from impinging Cretaceous Pacific swell –
therein) these chemostratigraphic pinning points provide a solid funda- is proposed (Fig. 2). A potential analogue, albeit not in size, is perhaps
ment for the stratigraphic framework applied here. found in the Tertiary Murray Basin in SE Australia (Lukasik et al.,
2000), whose environmental deposits pass gradually from a low energy
5.2. Carbon-isotope stratigraphy shallow subtidal environment to more open settings and, then into
outer ramp environment in sub-storm-wave environments. There,
The δ13Ccarb and δ13Corg data sets as described here were subdivided water masses were at least temporally sub-oxic in nature (compare
in chemostratigraphic segments (Pe1–Pe9; Fig. 9). In the case of the Inca with Fig. 6).
Formation, data are limited to δ13Corg due to the very low carbonate
content. 6.2. A carbon isotope reference curve for the sub-equatorial eastern Pacific
The δ13Corg section commences with an abrupt negative shift
reaching − 26.3‰ (Pe1), continued by a recovery phase (Pe2) and 6.2.1. Bulk-micrite carbon isotope ratios
followed by Pe3 that exhibits a gradual negative shift reaching −27‰. Secular changes in the isotopic composition of the dissolved
Further upsection, Pe4 reaches a recovery phase and Pe5 represents a inorganic carbon (DIC) pool of the world's oceans are, under favourable
plateau. The overlying Pe6 shows a sharp δ13Corg decrease down to conditions, recorded by marine carbonates (δ13Ccarb) resulting in char-
values as low as − 28‰. This feature is followed by a recovery phase acteristic chemostratigraphic patterns. This approach is widely accepted
(Pe7). Upsection, δ13Corg values exhibit a short negative shift (Pe8) and despite numerous problems and shortcomings, mid-Cretaceous
and finally a plateau phase (Pe9). OAEs are recognized and documented from bulk micrite δ13Ccarb curves
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 133

worldwide (Jarvis et al., 2002; Weissert and Erba, 2004; Erbacher et al., intervals of each section. The perhaps best test for the over-regional sig-
2005; Jenkyns, 2010; Wendler, 2013; Horikx et al., 2014). nificance of the isotope sections from Peru is their contrast-comparison
In order to provide a chemostratigraphic data set that has the poten- with well-established reference curves from other basins. Here, we
tial for supra-regional correlation, patterns in δ13Ccarb must at first be compare the Peru section with coeval successions in European Tethys
tested for diagenetic alteration masking palaeoceanographic patterns (Vocontian Basin, France; Piobbico section, Italy), the Pacific realm
(Marshall, 1992; Immenhauser et al., 2002). In shallow-neritic bulk (Western Platform, Peru; Pacific Guyots and Hokkaido, Japan) and the
micrite samples, meteoric waters related to transient subaerial expo- Western Atlantic domain (Sierra Madre, Mexico). Specific focus is
sure stages may introduce 18O-depleted oxygen, whereas 13C-depleted on mid-Cretaceous OAE signatures (OAEs 1b, 1c and 1d). In terms of
soil-zone CO2 can alter carbonate samples (Allan and Matthews, 1977; the temporal framework, this correlation is primarily based on the avail-
Christ et al., 2012). Generally, carbon isotope values are less prone to able biostratigraphic and chemostratigraphic analysis as documented
diagenesis as the carbonates form the dominant C reservoir compared here. Following previous well-established approaches, we assign
to oxygen ratios, sourced mainly from fluids (e.g., Marshall, 1992; chemostratigraphic labels to characteristic segments of the isotope
Jarvis et al., 2002). Moreover, fractionation of δ13Ccarb during burial is al- curves (Pe1 though Pe9; Fig. 12).
most insensitive to temperature-controlled processes (about 0.035‰ The isotope excursion Pe1 in the δ13Corg curve from Peru is dated by
per °C; Lynch-Stieglitz, 2003). In this context, it is of importance that the ammonite N. nicholsoni as early Albian in age (Benavides-Caceres,
we found no clear field evidence such as karsting, bleaching or traces 1956; Robert et al., 2009). The Pe1 event is correlated with the Kilian
of the roots of land plants in any of our sections. Similarly, evidence Level in the European Tethys (i.e., the second organic-rich level of the
for the typical saw-tooth-shaped isotope patterns described in Allan OAE1b bundle; Herrle et al., 2004; Reichelt, 2005; Fig. 12). In the
and Matthews (1977) and many subsequent papers is lacking. Similar Vocontian Basin, this interval consists of black, organic-rich sediment
to Jurassic ramps settings in Morocco (Christ et al., 2012), it is tentative- located close to the Aptian–Albian boundary (Bréhéret, 1986). This
ly concluded that low-amplitude sea-level fall was insufficient to expose level is tentatively correlated with the Monte Nerone black shale inter-
portions of the carbonate ramp studied here. Reasons for this might in- val in the Umbria Marche Basin, Italy (Erbacher, 1994).
clude rapid basement subsidence, a feature that is probably supported The second negative excursion Pe3, observed both in the δ13Ccarb and
13
by the generally stratigraphically thick successions studied. δ Corg records, is correlated with the lower Albian Paquier Level from
Given that the chemostratigraphy presented here is mainly based on the European Tethys (Fig. 12; Herrle et al., 2003, 2004; Reichelt,
bulk micrite data, the δ18O record of the analysed carbonates is not 2005). This is in line with the occurrence of Glottoceras raumundi within
interpreted in terms of its palaeoenvironmental significance but serves the Chulec Formation (Robert, 2002). The Paquier Level is associated
to understand the degree of diagenetic alteration of sections studied. with one of the most widespread carbon-isotope events of the mid-Cre-
Specifically, when plotting the carbon versus oxygen isotope data taceous and typically characterized by marine black shale located in the
from the Huameripashga and Pulluicana sections no correlation is ob- middle of the Hedbergella planispira foraminifera Zone of the Vocontian
served (R2 = 0.03) between the two. Often, but not always, the lack in Basin (Herrle et al., 2004). Furthermore, in the Western Atlantic, the
covariance between carbon and oxygen isotopes serves as an indication Paquier Level is recorded by organic-rich sediments belonging to the
of less than pervasive diagenetic resetting. Generally, the δ13Ccarb ratios H. planispira foraminifera Zone (Bralower et al., 1999). In terms of its
found are typical for mid-Cretaceous oceanic signatures reported from δ13C amplitude (N 1‰; Fig. 12), Pe3 in Peru is comparable with the
many basins worldwide (Veizer et al., 1999; Prokoph et al., 2008; mid-Pacific record (Jenkyns and Wilson, 1999).
Schulte et al., 2011). An exception is found in the significantly more neg- In segment Pe5, a 87Sr/86Sr ratio of 0.707380 ± 0.000006 was obtain-
ative ratios in the lower Yumagual Formation. This feature is not related ed from a well-preserved oyster shell (Fig. 9). This value is typical for an
to changes in facies or subaerial exposure-related diagenesis and might early to middle Albian, post Paquier Level, marine strontium isotope sig-
represent a local palaeoceanographic pattern that awaits further study. nature (Kennedy et al., 2000) and is in good agreement with our inter-
Summing up, whilst we do not exclude minor degrees of diagenetic al- pretation of the carbon isotope stratigraphy. A good correlation of
teration, the data shown here support the concept of a carbon isotope segment Pe5 in Peru is found with the Mexican record of Bralower
curve that predominantly reflects patterns in oceanic DIC. et al. (1999). The subsequent negative excursion Pe6 recorded in the
Pariatambo Formation is assigned to the middle Albian P. ulrichi ammo-
6.2.2. Bulk organic carbon isotope ratios nite Zone (Robert, 2002). Segment Pe6 occurs on top of a marine trans-
Patterns and amplitudes of δ13Corg ratios may result from variations gressive surface (SB1; Fig. 9) and spans the transgressive interval of the
in the source of organic carbon as well as from selective preservation MsS2. Segment Pe6 is tentatively correlated with the Leenhardt Level
and degradation of organic compounds during diagenesis (Bralower recorded in the European Tethys (top of the OAE1b set; Herrle et al.,
et al., 1999; Wendler, 2013; Trefry et al., 2014). In order to test the sig- 2004). The Leenhardt Level in the Vocontian Basin is characterized by
nificance of the δ13Corg curve obtained here we compare the pattern ob- a prominent, basin wide black shale interval (Bréhéret, 1986) in the
tained with that from inorganic carbon isotopes (Fig. 9). The observed planktonic foraminifer Ticinella primula Zone (Reichelt, 2005). Interest-
similarities are considered encouraging whilst the two records also ingly, in the Vocontian Basin, both the Paquier and the Leenhardt levels
show distinct differences in the amplitude of carbon isotope excursions. have been linked to transgressive phases (Bréhéret, 1994). A similar sit-
These may be related to variations in photosynthetic fractionation, dif- uation is observed in the Peruvian record. There, the two negative shifts
ferential organic carbon preservation (e.g., Jenkyns, 2010; Wendler, (Pe3 and Pe6) fall in the initial transgressive phase of MsS1 and 2
2013), differential sources of carbonate mud, differential diagenesis, fa- (Fig. 9). This may evidence widespread flooding and associated carbon
cies change or hidden hiatal surfaces (Immenhauser et al., 2008 and ref- cycle perturbations both for the Paquier and Leenhardt levels.
erences therein; Turpin et al., 2014). A prominent example is the The Pe8 negative shift in the Peruvian δ13Ccarb and δ13Corg section
isotope segment Pe8 (Fig. 9) that has a high-amplitude in δ13Ccarb, (Fig. 9) is also observed in limestones deposited during the Biticinella
whilst we observe a low-amplitude negative inflection in the δ13Corg breggiensis Zone in Mexico, on Pacific Guyots and in Japan (Bralower
curve (~ 1‰). In conclusion, the comparison between organic and et al., 1999; Jenkyns and Wilson, 1999; Takashima et al., 2004). In the
carbonate carbon isotopes supports the overall concept of a mainly Piobbico area in Italy, Pe8 occurs in the planktonic foraminifer Ticinella
palaeoceanographic record found. praeticinensis Subzone (Erbacher et al., 1996), but displays a subdued
amplitude only. In the northern Tethyan domain, Pe8 is best correlated
6.2.3. Comparison with other reference curves for Albian OAEs to a δ13Ccarb negative shift observed in the T. primula Zone, making this
The Pulluicana–Huameripashga composite δ13Ccarb and δ13Corg feature somewhat older than the other records. This discrepancy might
chemostratigraphy results from the stratigraphically most complete be explained by the error bars in age models applied. Above this
134
Eastern Pacific Central Pacific Western Pacific Western Atlantic Northern Tethys Western Tethys
13
Ccarb (‰ V-PDB)
13 13
Ccarb (‰ V-PDB) Corg (‰ V-PDB) Vocontian Basin 13
France Ccarb (‰ V-PDB)

J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137


Western Platform, Peru
(this study) (Reichelt, 2005) Piobbico area,
-3 -2 -1 0 1 2 3 -28 -27 -26 -25 1 2 3 Italy
(Erbacher et al., 1996)
CENOMANIAN

13

C
e
early

Corg (‰ V-PDB) 1 2 3 4

CENOMANIAN
Pe9b
Pe9b Hokkaido,
13
Japan Corg (‰ V-PDB)
(Takashima et al., 2004) Sierra Madre,
Breistoffer Level Pe9
? -26 -24 -24 Mexico

late
Pe9b
late

13
Ccarb (‰ V-PDB) (Bralower et al., 1999) OAE1d

ALBIAN
Pe9a Pacific Guyot -26 -25 -24 Pe9a
Pe9b

late
Amadeus (Jenkyns and Wilson, 1999) Pe9a

late
?

ALBIAN
Level

late
1 2 3 Pe9a Pe9 ?

ALBIAN
OAE1c?
Pe9
Pe8

middle
Pe8 Pe8
ALBIAN

middle
Pe8 Pe8 Pe7
Pe8

middle
Pe8 Pe7 Pe7
middle

Pe7? Pe6
ALBIAN

Pe7 ? ?
Pe7 Pe7 Pe6
500m Pe6 Pe4

early
middle
Leenhardt Pe4

ALBIAN
Pe3
Level ? Pe2 Pe3 OAE1b

l e
Pe6 Pe6 Pe6 set
Pe1

AP

AP
Pe5 40m 10m

l
Pe4

50m Pe5 Pe5 Pe3


20m
APTIAN

100m
early
early

Pe4 Pe4
Paquier level Pe4 Pe3
Pe3
Pe3 Pe2
AP

Pe1
l

Kilian Level

Fig. 12. Correlation of Peruvian δ13C reference curve for the subequatorial western Pacific with reference curves from the Central Pacific (Jenkyns and Wilson, 1999), the Western Pacific (Takashima et al., 2010), the Western Atlantic (Bralower et al.,
1999), the Northern Tethys (Reichelt, 2005) and the Western Tethys (Erbacher et al., 1996). Black shale intervals are shaded grey.
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 135

negative trend, the Amadeus Level (OAE1c, Erbacher et al., 1996) is co- the most prominent expression of the OAE1b event is assigned to the
eval to a short-lived, small-amplitude δ13C excursion to lower values. Jacob sub-event associated with the final demise of the neritic carbonate
This level, recorded in the Umbria Marche Basin was assigned to the factory (Föllmi et al., 2006), whereas in the Western Platform of Peru
B. breggiensis Zone and the T. praeticinensis Subzone (Central, Italy; the OAE1b bundle is best recorded by its final sub-event (Leenhardt
Galeotti, 1998). In summary, the contrast-comparison of the Peruvian Level).
composite isotope curve with coeval sections elsewhere is indicative Following the drowning related to the Leenhardt Level, the Peruvian
of a good comparability of Pacific and Western Atlantic records and in- sections reflect condensed sedimentation typified by a heterozoan
dicates that the sections measured recorded, and preserved, an above- biofacies in dark-grey mudstones. These condensed beds represent
regional chemostratigraphy. maximum flooding stages during LsS1 and MsS2 (Fig. 9). Deposition
In the Yumagual Formation, a positive excursion labelled Pe9a was probably influenced by ingression of more basinal, nutrient-rich
(Fig. 12) is present in the δ13Ccarb curve whereas it is probably absent waters from the Pacific during rising sea level (Trabucho-Alexandre
in organic carbon. In Europe, OAE1d is defined by a positive excursion et al., 2010). This oceanographic reorganization must have resulted in
(the Breistoffer Level; Bréhéret, 1997), characterized by black shale ho- reduced carbonate production and episodic shutdown of ramp carbon-
rizons dated as Rotalipora appenninica Zone (Reichelt, 2005). In Peru, ate deposition. This concept is in line with a shift in δ13Ccarb and δ13Corg
the δ13Ccarb Pe9a segment corresponds to that in the Japan curve towards more negative values (Pe8). Differences in amplitude may re-
(Pe9a). The observed partial discrepancy in Peruvian curves is interest- flect local environmental patterns in the partitioning of carbon between
ing but not uncommon in chemostratigraphy. Interpretations common- organic carbon and carbonate or variations in the values of total dis-
ly brought forward include local patterns in organic carbon cycling or solved inorganic carbon. Interestingly, the Pe8 event seems to be most
secondary features such as differential preservation of organic matter pronounced in the Pacific sections (including Japan) and in Mexico,
or differential diagenesis of δ13Ccarb. whereas Tethys sections do not reveal this event. This may point to a
circum-Pacific pattern not affecting the Tethyan realm (Fig. 12).
6.3. Impact of environmental changes on Peruvian neritic carbonate factory During the Albian–Cenomanian transition, an oyster-rich carbonate
ramp is established. Carbonate production and sedimentation rates
The carbonate–siliciclastic ramp deposits in Peru document tempo- are high. Here, both the δ13Ccarb and δ13Corg chemostratigraphy exhibit
ral changes in sedimentary environments and relative sea-level fluctua- a plateau phase, suggesting stable sedimentation, oceanographic condi-
tions. Moreover, observed patterns of enhanced carbonate production tions and a phase of regional tectonic quiescence.
or demise probably relate to the impact of transient carbon cycle pertur-
bations (OAEs 1b, 1c and 1d). 7. Conclusions
Within the OAE1b set, the negative excursion of the Kilian Level is
recorded in the North Atlantic and in the Western Tethys. Between An Albian–Cenomanian composite section from the vast Western
sections, considerable similarities include the coeval occurrence of Platform in Northern Peru comprises a total of seven facies associations.
organic-rich facies (Bralower et al., 1999; Herrle et al., 2004; These are assigned to a shallow to distal ramp environment with the
Trabucho-Alexandre et al., 2011). In the Peruvian sections, the Pe1 most proximal portions not being exposed in the study area.
δ13Corg negative excursion is probably the equivalent of the Kilian In the shallow subtidal inner ramp setting, argillaceous deposits are
Level (Fig. 9). In Peru, this event is recorded by marls and sandstones al- intercalated with restricted muddy carbonates. The open mid-ramp is
ternating with iron-rich facies of a shallow subtidal environment. These mainly characterized by oyster biostromes established beneath the ef-
are clearly different lithological patterns when compared to the Tethyan fective fair-weather wave base. Storms eroded, entrained and re-
and North Atlantic realms. Along the northern Tethyan margin (Helvetic deposited oyster debris resulting in widespread oyster packstone to
Platform), the onset of the OAE1b set is associated with the final demise rudstone facies. The outer ramp setting is located below the reach of
of the neritic carbonate factory (Föllmi et al., 2006) giving way to the effective storm wave base. Owing to the potential barrier effects of
siliciclastic sedimentation and phosphogenesis. Due to a gap in expo- the Albian volcanic arc to the west (Marañon Massif) and wave degra-
sure in the Peruvian sections, the recovery of the negative δ13Corg excur- dation across the shallow ramp, wave-base depths were probably
sion is not recorded. The Pe2 excursion might, however, represent an shallow.
increase in terrestrial organic matter influx during the turnover phase Both, the Marañon Massif and the Brazilian Shield to the east of the
from a mainly argillaceous sedimentation to a neritic carbonate factory. Western Platform represent important source areas for continental run-
This pattern is in good agreement with the negative δ13C excursion of off whilst currents transported proto-Atlantic and Pacific nutrient-rich
the Kilian Level, showing the characteristics of a global carbon cycle per- water masses onto to the Western Platform as expressed in a tropical
turbation (Jenkyns and Wilson, 1999). heterozoan carbonate factory. The Western Platform in Peru recorded
The expression of the Paquier Level in Peru (Pe3 segment) is, in characteristic global patterns in Albian and early Cenomanian δ13Ccarb
terms of its facies, represented by an alternation of wackestones and and δ13Corg ratios. The carbon isotope stratigraphy is constrained by bio-
dark-grey mudstones. Here, the base of the Pe3 segment corresponds stratigraphic data and by two oyster-based 87Sr/86Sr marker points. The
to a change from a heterozoan-dominated carbonate factory to a de- correlation with other reference sections in the Central and Western Pa-
creasing benthic carbonate production in the lower portion of the cific, the Western Atlantic and the Tethyan realm evidences that the
MsS1 transgressive phase (Fig. 9). In our view, this shift in sedimenta- Peru chemostratigraphy is of over-regional significance and has poten-
tion represents an incipient phase of platform demise indicative of a tial as a new reference curve for the subequatorial eastern Pacific realm.
global perturbation of the carbon cycle during the Paquier Level. The ab- Important chemostratigraphic features observed include the OAE1b set
sence of organic-rich deposits in the Pacific realm (Peru and mid-Pacific (Kilian, Paquier and Leenhardt levels), OAE1c and OAE1d.
Guyot) indicates that the accumulation of organic matter was either due The negative excursion of the Kilian Level in the Peruvian data,
to local signatures in carbon cycling superimposed on global patterns representing part of the OAE1b set, is probably representative of a global
(Robinson et al., 2004) or due to the deposition (or preservation) of or- carbon cycle perturbation and linked to the transition from a
ganic matter preferentially in deeper settings. siliciclastic-dominated sedimentation to a neritic carbonate factory.
In Peru, the isotopic signature of the Leenhardt level (Pe6; Fig. 12) is The expression of the Paquier Level corresponds to a change from a
found in dark-grey mudstones. These probably correspond to a major heterozoan-dominated carbonate factory to decreasing benthic carbon-
demise of neritic carbonate production marked in the field by a major ate production during incipient carbonate factory deterioration. The im-
discontinuity surface associated to the MsS2 transgressive surface pact of the Leenhardt level in Peru is characterized by major demise of
(SB1, Fig. 9). Conversely, in the Tethyan and Atlantic deep-water record, neritic carbonate production, marked in outcrops by a regional
136 J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137

discontinuity. This surface is followed by condensed sedimentation, Emeis, K.C., Weissert, H., 2009. Tethyan–Mediterranean organic carbon-rich sediments
from Mesozoic black shales to sapropels. Sedimentology 56, 247–266.
triggered by the influx of more basinal, nutrient-rich water masses Erbacher, J., 1994. Entwicklung und Paläoozeanographie mittelkretazischer Radiolarien
from the Palaeo-Pacific. Given that a similar pattern is not known der westlichen Tethys (Italien) und des Nordatlantiks. Tübinger Mikropaläontologie
from the Tethyan realm, this feature seems circum-Pacific in nature. Mitteilungen 12, 1–120.
Erbacher, J., Thurow, J., Littke, R., 1996. Evolution patterns of radiolaria and organic matter
Further upsection, an Albian-Cenomanian transition oyster-rich carbon- variations: a new approach to identify sea-level changes in mid-Cretaceous pelagic
ate ramp is established, suggesting stable environmental conditions. environments. Geology 24, 499–502.
The reference curve shown here is significant as (i) the vast area of Erbacher, J., Friedrich, O., Wilson, P.A., Birch, H., Mutterlose, J., 2005. Stable organic carbon
isotope stratigraphy across Oceanic Anoxic Event 2 of Demerara Rise, western tropi-
the subequatorial eastern Pacific has previously not been covered in cal Atlantic. Geochem. Geophys. Geosyst. 6, Q06010.
terms of its mid-Cretaceous carbon-isotope record and (ii) pronounced Föllmi, K.B., 2012. Early Cretaceous life, climate and anoxia. Cretac. Res. 35, 230–257.
similarities with well-established chemostratigraphic patterns else- Föllmi, K.B., Godet, A., Bodin, S., Linder, P., 2006. Interactions between environmental
change and shallow-water carbonate build-up along the northern Tethyan margin
where evidence the over-regional significance of the Peruvian data.
and their impact on the early Cretaceous carbon-isotope record. Paleoceanography
Nevertheless, the enormous dimensions of the Western Platform re- 21, PA4211.
quire the incorporation of regionally distributed coeval sections in Gale, A.S., Kennedy, W.J., Burnett, J.A., Caron, M., Kidd, B.E., 1996. The Late Albian to Early
order to shed light on spatial variations in carbon isotope stratigraphy. Cenomanian succession at Mont Risou near Rosans (Drôme, SE France): an integrated
study (ammonites, inoceramids, planktonic foraminifera, nannofossils, oxygen and
carbon isotopes). Cretac. Res. 17, 515–606.
Acknowledgements Gale, A.S., Bown, P., Caron, M., Crampton, J., Crowhurst, S.J., Kennedy, W.J., Petrizzio, M.R.,
Wray, D.S., 2011. The uppermost middle and upper Albian succession at the Col de
Palluel, Hautes-Alpes, France: an integrated study (ammonites, inoceramid bivalves,
This project was supported by the Deutsche Forschungsgemeinschaft planktonic foraminifera, nannofossils, geochemistry, stable oxygen and carbon iso-
(DFG, project no. BO-365/2-1) and by the Deutscher Akademischer topes, cyclostratigraphy). Cretac. Res. 32, 59–130.
Galeotti, S., 1998. Planktic and benthic foraminiferal distribution patterns as a response to
Austauschdienst (DAAD) through a scholarship to J.P.N. (PKZ:
changes in surface fertility and ocean circulation: a case study from the Late Albian
A/11/97387). We thank the Geological Survey of Peru (INGEMMET) for “Amadeus Segment” (Central Italy). J. Micropalaeontol. 17, 87–96.
important logistical support. W. Gomez is thanked for his help during Halfar, J., Godinez-Orta, L., Mutti, M., Valdez-Holguin, J.E., Borges, J.M., 2004. Nutrient and
temperature controls on modern carbonate production: an example from the Gulf of
field expedition. Analytical work was performed in the isotope laborato-
California, Mexico. Geology 32, 213–216.
ries at Bochum and Hannover by A. Niedermayr, D. Buhl and C. Wenske. Hallock, P., Schlager, W., 1986. Nutrient excess and the demise of coral reefs and carbon-
The present manuscript benefitted from the comments of two anony- ate platforms. Palaios 1, 389–398.
mous Palaeo3 reviewers and editorial guidance by F. Surlyk. Hay, W.W., Floegel, S., 2012. New thoughts about the Cretaceous climate and oceans.
Earth Sci. Rev. 115, 262–272.
Heimhofer, U., Hochuli, P.A., Herrle, J.O., Andersen, N., Weissert, H., 2004. Absence of
Appendix A. Supplementary data major vegetation and palaeoatmospheric pCO2 changes associated with oceanic an-
oxic event 1a (Early Aptian, SE France). Earth Planet. Sci. Lett. 223, 303–318.
Heimhofer, U., Hochuli, P.A., Herrle, J.O., Weissert, H., 2006. Contrasting origins of Early
Supplementary data to this article can be found online at http://dx. Cretaceous black shales in the Vocontian basin: evidence from palynological and cal-
doi.org/10.1016/j.palaeo.2015.01.025. careous nannofossil records. Palaeogeogr. Palaeoclimatol. Palaeoecol. 23, 93–109.
Herrle, J.O., Pross, J., Friedrich, O., Kossler, P., Hemleben, C., 2003. Forcing mechanisms for
mid-Cretaceous black shale formation: evidence from the upper Aptian and lower
References Albian of the Vocontian Basin (SE France). Palaeogeogr. Palaeoclimatol. Palaeoecol.
190, 399–426.
Allan, J.R., Matthews, R.K., 1977. Carbon and oxygen isotopes as diagenetic and strati- Herrle, J.O., Kossler, P., Friedrich, O., Erlenkeuser, H., Hemleben, C., 2004. High-resolution
graphic tools: surface and subsurface data, Barbados, West Indies. Geology 5, 16–20. carbon isotope records of the Aptian to lower Albian from SE France and the Mazagan
Ando, A., Kaiho, K., Kawahata, H., Kakegawa, K., 2008. Timing and magnitude of early Plateau (DSDP Site 545): a stratigraphic tool for palaeoceanographic and
Aptian extreme warming: unraveling primary δ18O variation in indurated pelagic car- palaeobiologic reconstruction. Earth Planet. Sci. Lett. 218, 149–161.
bonates at Deep Sea Drilling Project Site 463, central Pacific Ocean. Palaeogeogr. Horikx, M., Heimhofer, U., Dinis, J., Huck, S., 2014. Integrated stratigraphy of shallow ma-
Palaeoclimatol. Palaeoecol. 260, 463–476. rine Albian strata from the southern Lusitanian Basin of Portugal. Newsl. Stratigr.
Atherton, M.P., Aguirre, L., 1992. Thermal and geotectonic setting of Cretaceous volcanic http://dx.doi.org/10.1127/0078-0421/2014/0041.
rocks near Ica, Peru, in relation to Andean crustal thinning. J. South Amer. Earth Sci. Howarth, R.J., McArthur, J.M., 1997. Statistics for strontium isotope stratigraphy: a robust
5, 47–69. LOWESS fit to the marine Sr-isotope curve for 0 to 206 Ma, with look-up table for der-
Atherton, M.P., Webb, S., 1989. Volcanic facies, structure, and geochemistry of the margin- ivation of numeric age. Geology 105, 441–456.
al basin rocks of central Peru. J. South Amer. Earth Sci. 2, 241–261. Huck, S., Heimhofer, U., Rameil, N., Bodin, S., Immenhauser, A., 2011. Strontium and carbon-
Benavides-Caceres, V.E., 1956. Cretaceous system in Northern Peru. Am. Mus. Nat. Hist. isotope chronostratigraphy of Barremian–Aptian shoal-water carbonates: Northern
Bull. 108, 353–494. Tethyan platform drowning predates OAE 1a. Earth Planet. Sci. Lett. 304, 547–558.
Bodin, S., Godet, A., Westermann, S., Föllmi, K.B., 2013. Secular change in northwestern Immenhauser, A., 2005. High-rate sea-level change during the Mesozoic: new approaches
Tethyan water-mass oxygenation during the late Hauterivian–early Aptian. Earth to an old problem. Sediment. Geol. 175, 277–296.
Planet. Sci. Lett. 374, 121–131. Immenhauser, A., 2009. Estimating palaeo-water depth from the physical rock record.
Bralower, T.J., Cobabe, E., Clement, B., Sliter, W.V., Osburn, C.L., Longoria, J., 1999. The re- Earth Sci. Rev. 96, 107–139.
cord of global change in mid-Cretaceous (Barremian–Albian) sections from the Sierra Immenhauser, A., Kenter, J.A.M., Ganssen, G., Bahamonde, J.R., van Vliet, A., Saher, M.H.,
Madre, northeastern Mexico. J. Foraminifer. Res. 4, 418–437. 2002. Origin and significance of isotope shifts in Pennsylvanian carbonates (Asturias,
Bréhéret, J.G., 1986. Indices d'un énénement anoxique étendu à la Téthys alpine, à l'Albien NW Spain). J. Sediment. Res. 72, 82–94.
inférieur énénement Paquier. C. R. Acad. Sci. 300, 355–358. Immenhauser, A., Hillgärtner, H., Sattler, U., Bertotti, G., Schoepfer, P., Homewood, P.,
Bréhéret, J.G., 1994. The mid-Cretaceous organic-rich sediments from the Vocontian Zone Vahrenkamp, V., Steuber, T., Masse, J.P., Droste, H., Taal-van Koppen, J., Van der
of the French Southeast Basin. In: Mascle, A. (Ed.), Hydrocarbon and Petroleum Geol- Kooij, B., Van Bentum, E., Verwer, K., Hoogerduijn-Strating, E., Swinkels, W., Peters,
ogy of, France, pp. 295–320. J., Immenhauser-Potthast, I., Al Maskery, S., 2004. Barremian–lower Aptian Qishn For-
Bréhéret, J.G., 1997. L'Aptien et l'Albien de la Fosse Voconienne (bordures et basin): Évo- mation, Haushi–Huqf area, Oman: a new outcrop analogue for the Kharaib/Shu'aiba
lution de la sedimentation et enseignements sur les événements anoxiques. Soc. reservoirs. GeoArabia 9, 153–194.
Géol. Nord 25, 602. Immenhauser, A., Hillgärtner, H., van Bentum, E., 2005. Microbial–foraminiferal episodes
Christ, N., Immenhauser, A., Amour, F., Mutti, M., Preston, R., Whitaker, F., Peterhänsel, A., in the Early Aptian of the southern Tethyan margin: ecological significance and pos-
Egenhoff, A.O., Dunn, P.A., Agar, S.M., 2012. Triassic Latemar cycle tops — subaerial ex- sible relation to oceanic anoxic event 1a. Sedimentology 52, 77–99.
posure of platform carbonates under tropical arid climate. Sediment. Geol. 265–266, Immenhauser, A., Holmden, C., Patterson, W.P., 2008. Interpreting the carbon-isotope re-
1–29. cord of ancient shallow epeiric seas: lessons from the recent. In: Pratt, B.R., Holmden,
Coccioni, R., Galeotti, S., 1993. Orbitally induced cycles in benthonic foraminiferal C. (Eds.), Dynamics of Epeiric Seas, Special Publication vol. 48. Geological Association
morphogroups and trophic structures distribution patterns from the Late Albian of Canada, Toronto, Ontario, pp. 135–174.
‘Amadeus Segment’ (Central Italy). J. Micropalaeontol. 12, 227–239. INGEMMET, 1996. Regional Geological Map of Peru, scale 1: 1,000,000. Technical Report
D'Hondt, S., Arthur, M.A., 1996. Late Cretaceous oceans and the cool tropic paradox. Sci- INGEMMET (Geological Survey of Peru), Lima.
ence 271, 1838–1840. Jaillard, E., 1987. Sedimentary evolution of an active margin during middle and upper Cre-
Dumitrescu, M., Brassell, S.C., Schouten, S., Hopmans, E.C., Sinninghe Damsté, J.S., 2006. In- taceous times: the North Peruvian margin from Late Aptian up to Senonian. Geol.
stability in tropical Pacific sea-surface temperatures during the early Aptian. Geology Rundsch. 76, 677–697.
34, 833–836. Jarvis, I., Mabrouk, A., Moody, R.T.J., De Cabrera, S.C., 2002. Late Cretaceous (Campanian)
Dunham, R.J., 1962. Classification of carbonate rocks according to their depositional tex- carbon isotope events, sea-level change and correlation of the Tethyan and Boreal
ture. Am. Assoc. Pet. Geol. Bull. 1, 108–121. realms. Palaeogeogr. Palaeoclimatol. Palaeoecol. 188, 215–248.
J.P. Navarro-Ramirez et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 423 (2015) 122–137 137

Jarvis, I., Gale, A.S., Jenkyns, H.C., Pearce, M.A., 2006. Secular variation in Late Cretaceous foraminifera faunas. (Dissertation zur Erlangung des Grades eines Doktors der
carbon isotopes: a new δ13C carbonate reference curve for the Cenomanian– Naturwissenschaften.), Universität Tübingen (125 pp.).
Campanian (99.6–70.6 Ma). Geol. Mag. 143, 561–608. Robert, E., 2002. La transgression albienne dans le Bassin Andin (Pérou, Équateur).
Jarvis, I., Lignum, J.S., Gröcke, D.R., Jenkyns, H.C., Pearce, M.A., 2011. Black shale deposition, Paléontologie (ammonites), biostratigraphie et séqCuences de dépôts. Strata 38
atmospheric CO2 drawdown, and cooling during the Cenomanian–Turonian oceanic (380 pp.).
anoxic event. Paleoceanography 26, PA3201. Robert, E., Bulot, L.G., 2004. Origin, phylogeny, faunal composition, and stratigraphical sig-
Jenkyns, H.C., 2010. Geochemistry of oceanic anoxic events. Geochem. Geophys. Geosyst. nificance of the Albian engonoceratidae (pulchelliaceae, ammonitina) of Peru. J. S.
11, 1–30. Am. Earth Sci. 17, 11–23.
Jenkyns, H.C., Wilson, A., 1999. Stratigraphy, palaeoceanography, and evolution of Creta- Robert, E., Bulot, L.G., 2005. Albian ammonite faunas from Peru, the genus Neodeshayesites
ceous Pacific guyots: relics from a greenhouse Earth. Am. J. Sci. 299, 341–392. casey, 1964. J. Paleontol. 79, 611–618.
Kennedy, W.J., Gale, A.S., Bown, P.R., Caron, M., Davey, R.J., Gröcke, D., Wray, D.S., 2000. In- Robert, E., Latil, J.L., Bulot, L.G., 2009. Albian ammonite faunas from South America: the
tegrated stratigraphy across the Aptian–Albian boundary in the Marnes Bleues, at the genus Tegoceras hyatt, 1903. Rev. Paléobiol. 28, 43–51.
Col de Pré-Guittard, Arnayon (Drôme), and at Tartonne (Alpes-de-Haute-Provence), Robinson, S.A., Williams, T., Bown, P.R., 2004. Fluctuations in biosiliceous production and
France: a candidate Global Boundary Stratotype Section and Boundary Point for the the generation of Early Cretaceous oceanic anoxic events in the Pacific Ocean
base of the Albian Stage. Cretac. Res. 21, 591–720. (Shatsky Rise, Ocean Drilling Program Leg 198). Paleoceanography 19, PA4024.
Keulengan, G.H., Krumbein, W.C., 1949. Stable configuration of bottom slope in shallow Schlager, W., 2005. Carbonate sedimentology and sequence stratigraphy. Society Eco-
sea and its bearing on geological processes. Trans. Am. Geophys. Union 30, 855–861. nomic Paleontologists & Mineralogists, Concepts in Sedimentology and Palaeontology
Krencker, F.N., Bodin, S., Hoffmann, R., Suan, G., Mattioli, E., Kabiri, L., Föllmi, K.B., Series no. 8 (200 pp.).
Immenhauser, A., 2014. The middle Toarcian cold snap: trigger of mass extinction Schlanger, S.O., Jenkyns, H.C., 1976. Cretaceous oceanic anoxic events: causes and conse-
and carbonate factory demise. Glob. Planet. Chang. 117, 64–78. quences. Geol. Mijnb. 55, 179–184.
Kuypers, M.M.M., Blokker, P., Erbacher, J., Kinkel, H., Pancost, R.D., Schouten, S., Damste, Schröder-Adams, C.J., Herrle, J.O., Tu, Q., 2012. Albian to Santonian carbon isotope excur-
J.S.S., 2001. Massive expansion of marine Archaea during a mid-Cretaceous oceanic sions and faunal extinctions in the Canadian Western Interior Sea: recognition of eu-
anoxic event. Science 293, 92–94. static sea-level controls on a fore bulge setting. Sediment. Geol. 281, 50–58.
Larson, R.L., Pitman, W.C., 1972. World-wide correlation of Mesozoic magnetic anomalies, Schulte, P., van Geldern, R., Freitag, H., Karim, A., Négrel, P., Petelet-Giraud, E., Probst, J.L.,
and its implications. Geol. Soc. Am. Bull. 83, 3645–3662. Telmer, K., Veizer, J., Barth, J.A.C., 2011. Applications of stable water and carbon iso-
Leckie, R.M., Bralower, T.J., Cashman, R., 2002. Oceanic anoxic events and plankton topes in watershed research: weathering, carbon cycling, and water balances. Earth
evolution: biotic response to tectonic forcing during the mid-Cretaceous. Sci. Rev. 109, 20–31.
Paleoceanography 17, 1–29. Scott, R.W., Formolo, M., Rush, N., Owens, J.D., Oboh-Ikuenobe, F., 2013. Upper Albian
Lorenzen, J., Kuhnt, W., Holbourn, A., Flogel, S., Moullade, M., Tronchetti, Guy, 2013. A new OAE1d event in the Chihuahua Trough, New Mexico, U.S.A. Cretac. Res. 46, 136–150.
sediment core from the Bedoulian (lower Aptian) stratotype at Roquefort-La Bedoule, Soler, P., Bonhomme, M.G., 1990. Relations of magmatic activity to plate dynamics in
SE France. Cretac. Res. 39, 6–16. Central Peru from late Cretaceous to present. Geol. Soc. Am. Spec. Pap. 241, 173–192.
Luciani, V., Cobianchi, M., Fabbri, S., 2007. The regional record of Albian oceanic anoxic Sprovieri, M., Sabatino, N., Pelosi, N., Batenburg, S.J., Coccioni, R., Lavarone, M., Mazzola, S.,
events at the Apulian Platform Margin (Gargano Promontory, southern Italy). Rev. 2013. Late Cretaceous orbitally-paced carbon isotope stratigraphy from the
Micropaleontol. 50, 239–251. Bottaccione Gorge (Italy). Palaeogeogr. Palaeoclimatol. Palaeoecol. 379–380, 81–94.
Lukasik, J.L., James, N.P., McGowran, B., Bone, Y., 2000. An epeiric ramp: low-energy, cool- Takashima, R., Kawabe, F., Nishi, H., Moriya, K., Wani, R., Ando, H., 2004. Geology and stra-
water carbonate facies in a Tertiary inland sea, Murray Basin, South Australia. Sedi- tigraphy of forearc basin sediments in Hokkaido, Japan: Cretaceous environmental
mentology 47, 851–881. events on the Northwest Pacific margin. Cretac. Res. 25, 365–390.
Lynch-Stieglitz, J., 2003. Tracers of past ocean circulation. Treatise Geochem. 6, 433–451. Takashima, R., Nishi, H., Yamanaka, T., Hayashi, K., Waseda, A., Obuse, A., Tomosugi, T.,
Madhavaraju, J., IL Lee, Y., González-León, C.M., 2013. Diagenetic significance of carbon, Deguchi, N., Mochizuki, S., 2010. High-resolution terrestrial carbon isotope and
oxygen and strontium isotopic compositions in the Aptian–Albian Mural Formation planktic foraminiferal records of the Upper Cenomanian to the Lower Campanian in
in Cerro Pimas area, northern Sonora, Mexico. J. Iber. Geol. 39, 73–88. the Northwest Pacific. Earth Planet. Sci. Lett. 289, 570–582.
Marshall, J.D., 1992. Climatic and oceanographic isotopic signals from the carbonate rock Torsvik, T.H., Rousse, S., Labails, C., Smethurst, M.A., 2009. A new scheme for the opening
record and their preservation. Geol. Mag. 129, 143–160. of the South Atlantic Ocean and dissection of an Aptian Salt Basin. Geophys. J. Int. 177,
McArthur, J.M., Howarth, R.J., Bailey, T.R., 2001. Strontium isotope stratigraphy: LOWESS 1315–1333.
Version 3: best fit to the marine Sr-isotope curve for 0–509 Ma and accompanying Trabucho-Alexandre, J., Tuenter, E., Henstra, G.A., van der Zwan, K.J., van de Wal, R.S.W.,
look-up table for deriving numerical age. J. Geol. 109, 155–170. Dijkstra, H.A., de Boer, P.L., 2010. The mid-Cretaceous North Atlantic nutrient trap:
Melinte-Dobrinescu, M.C., Bojar, A.V., 2010. Late Cretaceous carbon- and oxygen isotope black shales and OAEs. Paleoceanography 25, PA4201.
stratigraphy, nannofossil events and palaeoclimate fluctuations in the Haţeg area Trabucho-Alexandre, J., Van Gilst, R.I., Rodríguez-López, J.P., de Boer, P.L., 2011. The sedi-
(SW Romania). Palaeogeogr. Palaeoclimatol. Palaeoecol. 293, 295–305. mentary expression of oceanic anoxic event 1b in the North Atlantic. Sedimentology
Moulin, M., Aslanian, D., Unternher, P., 2010. A new starting point for the South and Equa- 58, 1217–1246.
torial Atlantic Ocean. Earth Sci. Rev. 98, 1–37. Trabucho-Alexandre, J., Hay, W.W., de Boer, P.L., 2012. Phanerozoic environments of black
Mutterlose, J., Bodin, S., Fähnrich, L., 2014. Strontium-isotope stratigraphy of the Early shale deposition and the Wilson Cycle. Solid Earth 3, 29–42.
Cretaceous (Valanginian–Barremian): implications for Boreal–Tethys correlation Trefry, J.H., Trocine, R.P., Cooper, L.W., Dunton, K.H., 2014. Trace metals and organic car-
and palaeoclimate. Cretac. Res. 50, 252–263. bon in sediments of the northeastern Chukchi Sea. Deep Sea Res. Part II 102, 18–31.
Nederbragt, A.J., Fiorentino, A., Klosowska, B., 2001. Quantitative analysis of calcareous Turpin, M., Gressier, V., Bahamonde, J.R., Immenhauser, A., 2014. Component-specific pet-
microfossils across the Albian–Cenomanian boundary oceanic anoxic event at DSDP rographic and geochemical characterization of fine-grained carbonates along Carbon-
Site 547 (North Atlantic). Palaeogeogr. Palaeoclimatol. Palaeoecol. 166, 401–421. iferous and Jurassic platform-to-basin transects. Sediment. Geol. 300, 62–85.
Nemoto, T., Hasegawa, T., 2011. Submillennial resolution carbon isotope stratigraphy Veizer, J., Ala, D., Azmy, K., Bruckschen, P., Buhl, D., Bruhn, F., Carden, G.A.F., Diener, A.,
across the Oceanic Anoxic Event 2 horizon in the Tappu section, Hokkaido, Japan. Ebneth, S., Godderis, Y., Jasper, T., Korte, C., Pawellek, F., Podlaha, O.G., Strauss, H.,
Palaeogeogr. Palaeoclimatol. Palaeoecol. 309, 271–280. 1999. 87Sr/86Sr, δ13C and δ18O evolution of Phanerozoic seawater. Chem. Geol. 161,
Ogg, J.G., Hinnov, L.A., 2012. Chapter 27, Cretaceous. In: Gradstein, F.M., Ogg, J.G., Schmitz, 59–88.
M.D., Ogg, G.M. (Eds.), The Geological Time Scale. Elsevier, Amsterdam, pp. 793–853. Weissert, H., Erba, E., 2004. Volcanism, CO2 and palaeoclimate: a Late Jurassic–Early Cre-
Philip, J.M., Gari, J., 2005. Late Cretaceous heterozoan carbonates: Palaeoenvironmental taceous carbon and oxygen isotope record. J. Geol. Soc. Lond. 161, 695–702.
setting, relationships with rudist carbonates (Provence, south-east France). Sedi- Wendler, I., 2013. A critical evaluation of carbon isotope stratigraphy and biostratigraphic
ment. Geol. 175, 315–337. implications for Late Cretaceous global correlation. Earth Sci. Rev. 126, 116–146.
Pindell, J.L., Tabbutt, K.D., 1995. Mesozoic–Cenozoic Andean palaeogeography and region- Wilmsen, M., 2000. Evolution and demise of a mid-Cretaceous carbonate shelf: the Alta-
al controls on hydrocarbon systems. AAPG Mem. 62, 101–128. mira Limestones (Cenomanian) of northern Cantabria (Spain). Sediment. Geol. 133,
Price, G.D., 2003. New constraints upon isotope variation during the early Cretaceous 195–226.
(Barremian–Cenomanian) from the Pacific Ocean. Geol. Mag. 140, 513–522. Winter, L.S., Tosdal, R.M., Mortensen, J.K., Franklin, J.M., 2010. Volcanic stratigraphy and
Prokoph, A., Shields, G.A., Veizer, J., 2008. Compilation and time-series analysis of a ma- geochronology of the Cretaceous Lancones Basin, Northwestern Peru: position and
rine carbonate δ18O, δ13C, 87Sr/86Sr and δ34S database through Earth history. Earth timing of giant VMS deposits. Econ. Geol. 105, 713–743.
Sci. Rev. 87, 113–133. Yanin, B., 1983. Basics of Taphonomy: Moscow, Nedra (184 pp.).
Reichelt, K., 2005. Late Aptian–Albian of the Vocontian Basin (SE-France) and Albian of
NE-Texas: biostratigraphic and palaeoceanographic implications by planktic

You might also like