You are on page 1of 25

PUBLICATIONS

Paleoceanography
RESEARCH ARTICLE
10.1002/2014PA002772
Key Points:
Upper Aptian samples from the South
Atlantic Ocean are examined
Three dinocyst communities identied
in late Aptian
The oceanographic uctuations is
matched by the changes in dinocyst
communities

Correspondence to:
M. A. Carvalho,
mcarvalho@mn.ufrj.br

Citation:
Carvalho, M. A., P. Bengtson, and
C. C. Lana (2016), Late Aptian
(Cretaceous) paleoceanography of the
South Atlantic Ocean inferred from
dinocyst communities of the Sergipe
Basin, Brazil, Paleoceanography, 31,
226, doi:10.1002/2014PA002772.
Received 18 DEC 2014
Accepted 13 NOV 2015
Accepted article online 23 NOV 2015
Published online 9 JAN 2016

Late Aptian (Cretaceous) paleoceanography of the South


Atlantic Ocean inferred from dinocyst communities
of the Sergipe Basin, Brazil
Marcelo de A. Carvalho1, Peter Bengtson2, and Ceclia C. Lana3
1

Departamento de Geologia e Paleontologia, Museu Nacional, Universidade Federal do Rio de Janeiro, Rio de Janeiro, Brazil,
Institut fr Geowissenschaften, Universitt Heidelberg, Heidelberg, Germany, 3Gerncia de Bioestratigraa e Paleoecologia,
CENPES, PETROBRAS, Rio de Janeiro, Brazil

Abstract The late Aptian (Early Cretaceous) is a crucial time interval for understanding the paleoceanographic
changes in the Southern Hemisphere. Oceanographic changes in the emerging South Atlantic Ocean during this
interval are reected in the stratigraphic distribution of dinoagellate communities recorded in the Muribeca
and Riachuelo formations of the Sergipe Basin in northeastern Brazil. The Subtilisphaera community, in the lower
and middle parts of the section, appears to be related to the Subtilisphaera Ecozone and suggests the onset of
Tethyan inuence in the central South Atlantic, in a restricted to inner-neritic environment. The succeeding
Spiniferites community, in the middle part of the section, represents the rst signicant transgression, probably of
eustatic origin. The Cyclonephelium-Exochosphaeridium community, in the upper part of the section, appears to
be related to an oceanic event characterized by intermittent dysoxic-anoxic conditions. The uppermost part of
the section is dominated by the Spiniferites community, related to a progressive regional transgression and
culminating in an open-marine, fully Tethyan environment in the central part of the widening South Atlantic.
1. Introduction
Cretaceous marine transgressions were caused by some of the most extensive sea level rises during the
Phanerozoic. The oodings resulted in signicant changes in the depositional environments of the South
Atlantic continental-margin basins and may even have covered vast inland areas of the South American
continent during the mid-Cretaceous [Arai, 1999, 2007, 2009, 2014], creating seaways that connected the
Pacic Ocean with the North and South Atlantic oceans [Arai, 2009].
The Aptian transgression is recorded in carbonate platform deposits of tropical latitudes of South America (e.g.,
Venezuela and Colombia) and on the northern and southern margins of the Tethys, (e.g., Spain and France and
from Algeria to Oman) [Arnaud-Vanneau et al., 2008]. In Brazilian continental-margin basins the upper Aptian is
characterized by the rst major ooding surface observed in the Cretaceous succession. Data from paleontological studies [Azevedo, 2004; Arai, 2005, 2007, 2009, 2014] support the view that the mid-Cretaceous South
Atlantic was composed of a central and a southern part, separated by the Rio Grande RiseWalvis Ridge
(Figure 1). During the mid-Aptian to mid-Albian, the central South Atlantic, which contains the Sergipe Basin
(Figure 2), communicated mainly with the equatorial Atlantic (northern South Atlantic), which in turn was
connected to the Tethys [Bengtson et al., 2007; Koutsoukos and Bengtson, 2007]. However, the magnitude and
the precise dating of Tethyan inuence on the Brazilian continental-margin basins remain controversial.
Azevedo [2004] characterizes the circulation in the central South Atlantic as lagoonal; however, the presence
of foraminifers of ysch type, which are indicative of deeper-neritic to upper bathyal environments, suggests
the existence of deeper waters in northeastern Brazil [Koutsoukos, 1992].

2015. American Geophysical Union.


All Rights Reserved.

CARVALHO ET AL.

Musacchio [2000] and Arai [2009] suggested that although the Aptian transgression provided only restricted
marine connections, it caused pandemic distribution of selected paralic and continental organisms, for
example, ostracods and charophytes. Foraminifers and ammonites from the Sergipe Basin show conspicuous afnities with faunas of low latitudes of the Tethyan Realm [e.g., Bengtson and Koutsoukos,
1991, 1992; Koutsoukos et al., 1991; Koutsoukos, 1992; Koutsoukos and Bengtson, 1993; Bengtson and
Souza-Lima, 2000; Bengtson et al., 2007]. The global Aptian transgression resulted in the initial connection of the central South Atlantic with the equatorial Atlantic. The shift from the transitional, evaporitic
phase in the early Aptian to a fully marine drift phase was characterized by progressive crustal extension and thinning [Karner et al., 1992]. The gradual development of oceanic conditions in the central

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Figure 1. Idealized paleoceanographic setting for the late Aptian in the central and southern South Atlantic, after
Koutsoukos et al. [1991]; Azevedo [2004], and Arai [2014] showing location of the studied well GTP-24-SE. Reconstruction
at 119 Ma modied from ODSN Plate Tectonic Reconstruction Service. AA section = interpretation of a longitudinal
section extending from the equatorial region in the central to the southern South Atlantic.

South Atlantic established neritic-oceanic circulation patterns, and surface water was exchanged with
the equatorial Atlantic (equivalent to the South Atlantic and central Atlantic oceans, respectively, of
Bengtson and Koutsoukos [1992]), possibly even at intermediate epipelagic to mesopelagic water depths
[Koutsoukos, 1992].
Dinocysts (dinoagellate cysts) are frequently found in palynological samples from Cretaceous basins.
However, there are few previous reports of Cretaceous dinocyst assemblages from the Sergipe Basin. The
most relevant study is by Regali et al. [1974, 1975], who investigated several Brazilian basins, focusing mainly
on biostratigraphic applications. Other studies, carried out by geologists of the Brazilian oil company
Petrobras, remain unpublished.
According to several studies [Wall et al., 1977; Goodman, 1979, 1987; Harland, 1983; Brinkhuis and Zachariasse,
1988; Brinkhuis, 1992; Courtinat, 1993; Wilpshaar and Leereveld, 1994; Pearce et al., 2003; Lebedeva, 2008],
changes in physical and chemical features of water masses are reected in the distribution patterns of

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Figure 2. Location map of the continental-margin basins of northeastern Brazil [from Seeling, 1999].

modern dinocysts. Therefore, quantitative analyses of dinocyst assemblages may provide information
regarding depositional environment, sea level variations, sea surface temperature, productivity, and salinity.
Signicant progress in the development and application of dinocyst paleoecology has been made in the past
decades, especially where quantitative approaches are employed. Dinocyst analysis is a powerful tool in
paleoceanographic studies and particularly useful for linking inshore and offshore settings [Vandenberghe
et al., 2003].
Goodman [1979] formulated the concept of cyst communities, the lateral and vertical distributions of which
are attributed to the particular depositional environments. Thanks to the studies of Quaternary and Recent
cyst distribution, inferences can be made about pre-Quaternary dinocysts.
The present study aims at a better understanding of the origin and uctuation of the vertical dinocyst
distribution in the Sergipe Basin during the Aptian transgression, based on qualitative and quantitative
stratigraphic analyses of core samples. It highlights the importance of sea level and oceanic circulation in
controlling the distribution of dinocysts linked to the progressive separation of the African and South
American continents.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

2. Geological Context
The Sergipe Basin in northeastern Brazil contains one of the most extensive marine middle Cretaceous carbonate successions among the central South Atlantic basins. The basin, which forms the southern part of the
Sergipe-Alagoas Basin [Souza-Lima et al., 2002], is an elongate continental-margin basin situated between
latitudes 10151130S and longitudes 36203730W. Onshore the basin is 1650 km wide, 170 km long,
and covers an area of 6000 km2; the offshore portion comprises an area of ~5000 km2 (Figure 2).
Structurally, the Sergipe Basin forms a half-graben with a regional dip averaging 1015 to the southeast
[Ojeda and Fugita, 1976]. The basin is bounded by faults with an overall NE-SW and NW-SE orientation. The
faults were formed during rupture of the African-South American continent in the Early Cretaceous. The structural framework of the basin is characterized by large tilted fault blocks, which form structural lows and highs.
The Sergipe Basin belongs to the class of sedimentary basins characteristic of passive continental margins.
According to Ojeda and Fugita [1976] and Ojeda [1982], the evolution of the basin comprises ve main
tectono-sedimentary phases, viz., intracratonic, prerift, rift (earliest Cretaceous to early Aptian), transitional
(Aptian), and an open-marine drift phase (late Aptian to Recent). The upper Aptian succession is represented
by the upper part of the Muribeca Formation (Oiteirinhos Member) and the lower part of the Riachuelo
Formation (Taquari Member).

3. Studied Succession
The material for this study derives from well GTP-24-SE drilled by Petrobras/Petromisa in the TaquariVassouras area, between the towns of Rosrio do Catete and Carmpolis (Figure 2). The cored succession
has a thickness of 252.5 m, with samples taken at intervals of ~2.9 m. A total of 86 samples were analyzed.
The stratigraphic succession comprises parts of the Muribeca and Riachuelo formations (Figure 3a). In
general, the Muribeca Formation is composed of evaporites, clastics, and carbonate sediments laid down
during the transitional phase. The Riachuelo Formation was deposited during the open-marine phase and
consists of a carbonate-dominated sequence of calcareous mudstones and oolitic/oncolitic-bioclastic
grainstones/packstones, with subordinate conglomerates, sandstones, marls, and shales [Koutsoukos et al.,
1991]. In the studied samples, the Muribeca Formation is represented by the upper part of the Oiteirinhos
Member (21.5 m), which consists of intercalations of grey to black bituminous shales, limestones, and
siltstones. The Riachuelo Formation is represented by the lower part of the Taquari Member (231 m), composed
of alternating calcareous mudstones and shales. The lower half of the cored section (270125 m, lithofacies 1,
2a, and 2b) is dominated by calcareous mudstones and the upper half (12510 m, lithofacies 2c) by shales
(Figure 3b).
Analyses for foraminifers and calcareous nannofossils were carried out on all samples. Planktonic foraminifers
were found to be rare in the studied section. Samples analyzed between 249 m and 15 m revealed the
presence of indeterminate hedbergellids and few specimens of Favusella washitensis (Marta Claudia Viviers,
Petrobras, personal communication, 2015). These associations cannot be referred to global or local biozones
or specic ages. Nevertheless, similar faunas are commonly recovered from the upper Aptian of the Sergipe
Basin [Koutsoukos, 1989, 1992], corroborating the age inference based on palynological data (late Aptian,
P-270 biozone).
Available benthic foraminiferal data (samples from interval 27719 m) point to an overall shallow-neritic
environment with high terrigenous inux, the scarce microfauna being represented by agglutinant forms
and locally by the genus Lenticulina (interval 7942 m; Marta Claudia Viviers, Petrobras, personal communication, 2015). This calcareous-hyaline genus indicates the prevalence of dysoxic bottom-water conditions in
this interval.
Nannofossils were found in the section, but the analyses revealed no age-diagnostic species.
The samples yielded an abundant and diverse palynoora. The assemblages are typical of those reported
from upper Aptian strata of Brazilian continental-margin basins and can be referred to the Sergipea variverrucata Zone [Regali et al., 1975]. In the studied section, the uppermost occurrence of S. variverrucata is recorded
at 42.7 m. According to Regali and Santos [1999], in the Sergipe Basin the Sergipea variverrucata Zone is in part
correlated with the Ap-1 planktonic foraminiferal zone of Koutsoukos [1989], which in turn is correlated with

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Figure 3. (a) Geological cross-section of the Muribeca and Riachuelo formations of the Sergipe Basin [adapted from Borchert, 1977]. D = discontinuity according to
Mendes [1994]. Not to scale. (b) Lithostratigraphic scheme of well GTP-24-SE showing the gamma-ray curve and lithofacies [from Mendes, 1994]. Ts = transgressive
surface and mfs = maximum ooding surface [from Carvalho et al., 2006a].

the global upper Aptian Hedbergella infracretacea Zone and part of the upper Aptian Globigerinelloides algerianus
Zone [Koutsoukos and Bengtson, 1993] as presented in Figure 4. These correlations provide a late Aptian age for
the studied section, estimated at 118120(?) Ma.

4. Methods
Samples were prepared at the Research Centre of Petrobras (CENPES), Rio de Janeiro, using the standard
Petrobras method of palynological preparation compiled by Uesugui [1979] on the basis of methods developed
by Erdtman [1943, 1969] and Faegri and Iversen [1966], among others. In this method, all mineral constituents
are destroyed by hydrochloric and hydrouoric acids before heavy-liquid separation. The remaining organic
matter is sieved through a 10 m mesh and mounted on slides. The samples were analyzed under a transmitted
light microscope, and the analysis was based on the rst 200 palynomorphs counted on each slide.
The paleoenvironmental and paleoceanographic interpretation was based on the association of dinocysts
(dinoagellate cysts) revealed by cluster analysis to identify ecological similarities between palynomorph
communities from different depositional settings. Cluster analysis based on abundance and composition
was employed using the Ward method with Pearson-r similarity measure to establish groupings and recognize relationships between taxa. The cluster analysis forms discrete groupings based on abundances of the
objects. The results are displayed in dendrograms.
The Canonical Correspondence Analysis (CCA) [Legendre and Legendre, 1998] was performed using PAST software [Hammer et al., 2001]. This technique was chosen because the depths/species matrix is given for

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Figure 4. Biochronostratigraphic framework for the Aptian of the Sergipe Basin with the relationship between the Sergipea variverrucata pollen zone (shaded), local
planktonic foraminiferal zones (modied from Koutsoukos [1989]), global planktonic foraminiferal zones, and Tethyan ammonoid zones [from Ogg and Hinnov, 2012].

environmental variables, herein ecological parameters (equitability, amorphous organic matter/phytoclasts


ratio, productivity, and terrigenous input).
Diversity and equitability indices were calculated for all samples using the PAST software [Hammer et al.,
2001] to reconstruct paleoceanographic trends. The Shannon-Weaver diversity index [H(S)] takes into
account the abundance of each species to characterize the diversity of assemblages. Equitability was calculated to describe evenness within assemblages. A value of 1 represents equal abundances of all species,
whereas equitability approaches zero if one species dominates the assemblage.
The dominance for each sample, expressed as a percentage, was determined on the basis of the formula
D = (N1 + N2)/Nt, where N1 is the number of specimens of the most abundant species in a sample, N2 is
the number of specimens of the second most abundance species in that sample, and Nt is the total number
of specimens counted in the sample [Goodman, 1979].
The peridinioid to gonyaulacoid ratio (P/G) ratio, or heterotrophic/autotrophic ratio, introduced by Harland
[1973], reveals changes in primary productivity in the geologic past [e.g., Eshet et al., 1992; Versteegh, 1994;

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Brinkhuis et al., 1998]. A peridinioid-dominated assemblage indicates nutrient-rich and low-salinity conditions
related to freshwater inux into a marine environment [van Helmond et al., 2014]. By contrast, low values
of the ratio, i.e., gonyaulacoid-dominated assemblages, indicate open-marine environments. The ratio
used herein was described by Versteegh [1994] as P/G = nP/(nP + nG), where n is the number of specimens counted, P protoperidinioid dinoagellate cysts (P-cysts), and G gonyaulacoid dinoagellate
cysts (G-cysts).
The ratio of continental (terrestrial) to marine palynomorphs (C/M) was also described by Versteegh [1994],
calculated as C/M = nC/(nC + nM), where n is the number of specimens counted, C spores + pollen grains,
and M dinocysts + acritarchs. The C/M ratio was used herein to reconstruct changes in terrestrial input into
the basin. The same procedure was applied for the ratio of amorphous organic matter (AOM) and phytoclasts
(AOM/Phyto) to evaluate oxygen-depleted conditions in the environment, given that large amounts of AOM
result from environments with high preservation potential and low energy [Carvalho et al., 2013]. An
AOM-dominant percentage of total kerogen may be an indication of a reducing environment or at least a
temporarily dysoxic to anoxic environment with high preservation potential. In low-energy, proximal,
delta-top facies, some of the amorphous material present may be the product of degradation of higher plants
[Tyson, 1989, 1993, 1995; Mendona Filho et al., 2010; Pacton et al., 2011].
Total organic carbon (TOC) determinations and Rock-Eval pyrolysis data used herein are from Carvalho [2001]
(http://archiv.ub.uni-heidelberg.de/volltextserver/1586/1/marcelo.pdf) and Carvalho et al. [2006a], respectively.

5. Dinocysts
5.1. General Characteristics
Samples from the studied section yielded an abundant and diverse palynoora. All samples contain dinocysts
(Table 1), in most cases of moderate to high abundance, exceeding 100 individuals in ~50% of the slides
(Appendix). The samples differ in taxonomic composition, abundance, and preservation of the palynomorphs, with dinocysts moderately to well preserved.
Dinocysts are most abundant in the lower, mudstone-dominated part of the section (270125 m; Figure 3b)
but more diversied in the upper, shale-dominated part (12510 m; Figure 3b). In terms of abundance, no
signicant differences between the two lithologies are observed.
Twenty-nine species of dinocysts were identied in the samples (Table 1). The Shannon-Weaver diversity
index varies from 0.28 to 2.14, with an average of 1.19, and dominance varies from 0.17 to 0.95,
averaging 0.45.
The species Subtilisphaera senegalensis, Spiniferites chebca, and Spiniferites ramosus dominate the samples.
The rst species comprises between 0 and 93.4% of all individuals on a slide, averaging 30.9%, with
Spiniferites chebca and S. ramosus together reaching 30.3%.
The tectonic phases and lithologic characteristics shown in Figure 3 indicate a transgressive trend, from
transitional in the lower part to open marine in the upper part. The lower part of the section is dominated
by peridinioids of the Subtilisphaera community (Table 1 and Figure 6) coincident with the nearshore conditions pointed out previously and suggested by previous studies [Vozzhennikova, 1965; Scull et al., 1966;
Williams, 1977; Tappan, 1980; Sarjeant et al., 1987]. Higher in the section, where open-marine conditions
are inferred, gonyaulacoid forms are more common. In this part, the Oligosphaeridium community also occurs
and the Cyclonephelium-Exochosphaeridium community becomes predominant.
5.2. Dinocyst Communities
Four dinocyst communities are recognized in the upper Aptian of the Sergipe Basin, viz., the Oligosphaeridium,
Cyclonephelium-Exochosphaeridium, Spiniferites, and Subtilisphaera senegalensis communities. These were distinguished using cluster analysis (Figure 5) and named after their most dominant taxa (Figure 6). The cophenetic
correlation coefcient of 0.83 demonstrates the utility of the method.
5.2.1. Oligosphaeridium Community
The Oligosphaeridium community is the least abundant of the four dinocyst communities, represented by
only 1.5% of the total occurrences. It is dominated by Oligosphaeridium complex (56.8% of this community)
(Table 2) and also contains O. albertense, O. totum, O. poculum, O. pulcherrimum, and Systematophora spp.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Table 1. Summary of Species and Suggested Paleoenvironments Based on Specialized References


Species

References

Suggested Paleoenvironments
Nearshore, shallow-marine; inner neritic

Oligosphaeridium albertense
Oligosphaeridium complex
Oligosphaeridium pocolum
Oligosphaeridium pulcherrimum
Oligosphaeridium totum
Palaeoperidinium cretaceum

Millioud [1967], Lam and Porter [1977], Davey [1978], Burger


[1982], Hedlund and Norris [1986], Helby and McMinn [1992],
Duane [1994, 1996], Li and Habib [1996], and Skupien and
Vaek [2002]
Brinkhuis and Zachariasse [1988], Lister and Batten [1988],
Marshall and Batten [1988], Omran et al. [1990], Wilpshaar and
Leereveld [1994], Gedl [1999], Pearce et al. [2003], and
Mahmoud and Deaf [2007]
Downie et al. [1971], Brinkhuis and Zachariasse [1988], Lister
and Batten [1988], Marshall and Batten [1988], Omran et al.
[1990], Harker et al. [1990], Wilpshaar and Leereveld [1994],
Abdel-Kireem et al. [1996], Skupien and Vaek [2002], and
Mahmoud and Deaf [2007]
Liengjarern et al. [1980], Brinkhuis and Zachariasse [1988],
Marshall and Batten [1988], Courtinat and Schaaf [1990],
Harker et al. [1990], Omran et al. [1990], Eshet et al. [1992],
Abdel-Kireem et al. [1996], Li and Habib [1996], Lamolda and
Mao [1999], Harris and Tocher [2003], Lignum et al. [2007],
and Moustafa and Lashin [2012]
Goodman [1979] and Barroso-Barcenilla et al. [2011].
Li and Habib [1996], Lamolda and Mao [1999], and Peyrot et al.
[2012]
Omran et al. [1990]
Harker et al. [1990], Courtinat [1993], Wilpshaar and Leereveld
[1994], Abdel-Kireem et al. [1996], Gedl [1999]; Prauss [2001],
and Skupien and Vaek [2002]
Omran et al. [1990], Wilpshaar and Leereveld [1994], AbdelKireem et al. [1996], Duane [1996], Li and Habib [1996],
Lamolda and Mao [1999], Prauss [2001], Skupien and Vaek
[2002], Harris and Tocher [2003], and Mahmoud and Deaf
[2007]
Harding [1990], Prauss [2001], and Harris and Tocher [2003];

Pervosphaeridium spp.

Courtinat [1993], Harris and Tocher [2003], and Lebedeva [2008]

Pseudoceratium securigerum

Gedl [1999], Prauss [2001], Skupien and Vaek [2002],


Mahmoud and Deaf [2007], and Moustafa and Lashin [2012]
Downie et al. [1971], Davey and Rogers [1975], Brinkhuis and
Zachariasse [1988], Marshall and Batten [1988], Courtinat
and Schaaf [1990], Harker et al. [1990], Eshet et al. [1992],
Courtinat [1993], Abdel-Kireem et al. [1996], Lamolda and
Mao [1999], Prauss [2001], Skupien and Vaek [2002], and
Lignum et al. [2007]
Davey and Rogers [1975], Jain and Millepied [1975], Wall et al.
[1977], Regali [1989a], Omran et al. [1990], Arai et al. [1994],
Brinkhuis [1994], Powell et al. [1996], Gedl [1999], Rochon
et al. [1999], Arai et al. [2000], Antonioli [2001], Antonioli and
Arai [2002], Skupien and Vaek [2002], Harris and Tocher
[2003], Arai [2007], Mahmoud and Deaf [2007], and Moustafa
and Lashin [2012]
Brinkhuis and Zachariasse [1988], Lister and Batten [1988],
Marshall and Batten [1988], Omran et al. [1990], Wilpshaar
and Leereveld [1994], Mahmoud and Deaf [2007]
Brinkhuis et al. [1998] and Harris and Tocher [2003].
Courtinat and Schaaf [1990], Courtinat [1993], Abdel-Kireem
et al. [1996], Harris and Tocher [2003], and Peyrot et al. [2011]

Apteodinium granulatum

Circulodinium distinctum

Cribroperidinium edwardsii

Cyclonephelium vannophorum

Dinopterygium cladoides
Exochosphaeridium phragmites
Florentinia mantellii
Odontochitina operculata

Spiniferites ancorifer
Spiniferites bejui
Spiniferites chebca
Spiniferites lenzi
Spiniferites seghris
Spiniferites ramosus
Subtilisphaera cheit
Subtilisphaera perlucida
Subtilisphaera pirnaensis?
Subtilisphaera senegalensis

Systematophora cretacea

Tanyosphaeridium sp.
Trichodinium castanea

CARVALHO ET AL.

Brackish and littoral; marginal-marine (deltaic); inner-neritic


restricted marine

Most common in stable marine environments; inner-neritic


and restricted marine; marginal-marine (deltaic); innerneritic

Marginal, brackish, coastal; nearshore, unstable environments;


stressed environment; euryhaline

Relatively more inshore, marine restricted


Nearshore, shallow-marine, associated with high carbonate
content
Open-marine mid-shelf
Stable marine environments; normal salinity and oxygenation,
brackish and littoral, salty marshy (restricted shallowmarine); restricted shallow-marine, reduced salinity
Open-marine mid-shelf; euryhaline (O. pulcherrimum);
stenohaline (O. totum); open-marine (inner-neritic), neritic;
neritic to outer neritic; warmer and/or deeper shelf water

Euryhaline; brackish water; normal marine salinity to estuarine


conditions ; enhanced primary productivity, estuarine
circulation
No environmental preferences; normal salinity and
oxygenation; uniform habitat
Brackish and littoral; marginal-marine; marginal-marine
(deltaic); nearshore conditions
Normal salinity and oxygenation, mid-shelf, stable marine
environment; nearshore shallow-water environments and
fall-off in an offshore direction; neritic; warmer and/or
deeper shelf

Marginal, brackish, coastal; brackish and littoral in warmer


water; stenohaline (S. pirnaensis); marginal-marine (deltaic)
environment; salty marshy (restricted shallow-marine)

Inner-neritic and restricted-marine;, marginal-marine (deltaic)

No environmental preferences
Paleoenvironments with high content of TOC; normal salinity
and oxygenation; no environmental preferences; outerneritic preferences

LATE APTIAN SOUTH ATLANTIC, DINOCYST

Paleoceanography

10.1002/2014PA002772

Figure 5. Ward Dendrogram (r-mode) of 15 dinocyst genera from the studied section showing the four communities.

The Oligosphaeridium community contains species with long processes, a feature assumed to be an indication of open-marine, neritic conditions (Table 1). Also, Oligosphaeridium pulcherrimum is euryhaline or able
to adapt to a wide range of salinities [Harris and Tocher, 2003].
The Oligosphaeridium community occurs in only ~50% of the samples, mainly in the upper part of the section.
At 65.9 m, this community displays a pronounced peak representing 37.5% of the total communities (Figure 7).
It exceeds the general average of the community (1.5%) in only 15 samples. The community is rare in the lower
part of the section, which is dominated by the Subtilisphaera community (Figure 7).
5.2.2. Cyclonephelium-Exochosphaeridium Community
The Cyclonephelium-Exochosphaeridium community represents 31.6% of the dinocyst communities, making it the
second most abundant. This community is composed of Cyclonephelium vannophorum, Exochosphaeridium
phragmites, Circulodinium distinctum, Pervosphaeridium spp., Apteodinium granulatum, Tanyosphaeridium spp.,
Pseudoceratium securigerum, and Odontochitina operculata. Exochosphaeridium phragmites (46.4% of the community) and Cyclonephelium vannophorum (42.3% of the community) are the dominant species (Table 2). This
community is present in 76 of the 86 samples and dominates the middle part of the section, where it shows three
distinctive peaks, with the highest-abundance peak at 160.3 m (73.0%) (Figure 7). The community also shows two
pronounced peaks in the upper part of the section (Figure 7).
The relationship between the two most abundant genera, Cyclonephelium and Exochosphaeridium, has been
discussed by other authors, specically with respect to the Cenomanian-Turonian transition [Li and Habib,
1996; Harris and Tocher, 2003; Lignum et al., 2007; Peyrot et al., 2011; van Helmond et al., 2014].
Exochosphaeridium phragmites has been interpreted as typical of deeper water [Vozzhennikova, 1965; Scull
et al., 1966; Williams, 1977; Tappan, 1980; Sarjeant et al., 1987] but has also been associated with proximal environments and in some cases even very restricted environments [Peyrot et al., 2011]. The species of the

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

10

Paleoceanography

10.1002/2014PA002772

Figure 6. Dinocysts identied in the studied section: (A) Subtilisphaera senegalensis, (b) Spiniferites chebca, (c) Spiniferites
ramosus, (d) Cyclonephelium vannophorum, (e) Exochosphaeridium phragmites, (f) Oligosphaeridium complex. Scale bar = 20 m.

Cyclonephelium-Exochosphaeridium community are interpreted as indicative of shallow to bathyal conditions (Table 1). In some cases, Cyclonephelium is associated with more environmentally stressed conditions
[e.g., Eshet et al., 1992; Courtinat, 1993; Lana, 1997, 1998; Lignum et al., 2007]. Most of the species composing the community are indicative of a shallow-marine environment, exceptions being Pervosphaeridium
spp. and Tanyosphaeridium spp., which show very low abundances (0.1 and 0.3%, respectively) and no
environmental preferences [Harris and Tocher, 2003]. The species Circulodinium spp., Apteodinium granulatum, Pseudoceratium securigerum, and Odontochitina operculata make up 11% of the CyclonepheliumExochosphaeridium community and are all indicative of nearshore conditions (Table 1).
5.2.3. Spiniferites Community
The Spiniferites community is the most abundant community, comprising 38.2% of the total communities.
The dominant species are Spiniferites chebca and Spiniferites ramosus (together 79.6% of this community).

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

11

Paleoceanography

10.1002/2014PA002772

Table 2. Average Abundances (%) of Each Species and Its Respective Assemblages for the Studied Section
Communities
Oligosphaeridium community

Cyclonephelium-Exochosphaeridium community

Spiniferites community

Subtilisphaera community

Species

Oligosphaeridium complex
Oligosphaeridium albertense
Oligosphaeridium totum
Systematophora cretacea
Oligosphaeridium pulcherrimum
Oligosphaeridium pocolum
Exochosphaeridium phragmites
Cyclonephelium vannophorum
Pseudoceratium securigerum
Apteodinium granulatum
Circulodinium distinctum
Tanyosphaeridium spp.
Odontochitina operculata
Pervosphaeridium spp.
Spiniferites chebca
Spiniferites ramosus
Trichodinium castanea
Spiniferites bejui
Palaeoperidinium cretaceum
Dinopterygium cladoides
Florentina mantelli
Spiniferites ancorifer
Subtilisphaera senegalensis
Spiniferites lenzi
Spiniferites seghris
Subtilisphaera cheit
Subtilisphaera pirnaensis?
Subtilisphaera perlucida?
Cribroperidinium edwardsii
Total of Oligosphaeridium community
Total of Cyclonephelium-Exochosphaeridium community
Total of Spiniferites community
Total of Subtilisphaera community

56.8
25.3
7.4
4.2
4.2
2.1
46.4
42.3
5.3
4.4
1.0
0.3
0.3
0.1
41.0
38.6
17.4
1.4
0.9
0.5
0.2
0.04
92.9
5.6
0.6
0.4
0.3
0.2
0.1
1.5
31.6
38.2
28.6

The community also includes Trichodinium castanea, Palaeoperidinium cretaceum, Spiniferites ancorifer,
Florentinia mantellii, Spiniferites bejui, and Dinopterygium cladoides (Table 2).
Species of Spiniferites make up 81.0% of this community. The third most abundant species, Trichodinium
castanea, is only moderately abundant (averaging 17.4% of the community). According to several studies
(see Table 1), this species is indicative of normal salinity and oxygenation. Additionally, very low abundances
of two species indicative of inshore conditions, Dinopterygium cladoides and Palaeoperidinium cretaceum [e.g.,
Courtinat and Schaaf, 1990; Courtinat, 1993; Abdel-Kireem et al., 1996; Harris and Tocher, 2003; Peyrot et al.,
2011], are recorded in the community (together 1.4%).
The Spiniferites community is particularly dominant in intervals characterized by low abundances of the
Subtilisphaera community (Figure 7). The community is particularly abundant in the upper part of the
section, where it reaches 79.2% of all communities (at 20.85 m), and is absent from the lower part of
the section. However, the Spiniferites community shows conspicuous uctuations in abundance
(Figure 7).
5.2.4. Subtilisphaera Community
The Subtilisphaera community represents 28.6% of the total communities and is dominated by Subtilisphaera
senegalensis (92.9% of this community) (Table 2), which is the most abundant species in the whole section
studied. It contains Subtilisphaera cheit, Subtilisphaera pirnaensis?, Subtilisphaera perlucida?, Spiniferites lenzi,
Spiniferites seghris, and Cribroperidinium edwardsii.
The Subtilisphaera community occurs in all samples and is the only community present in the lower part of
the section (268.05 m) (Figure 7). Abundances show strong uctuations from the middle to the upper part
of the section, with lower values related to a high abundance of the Spiniferites community.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

12

Paleoceanography

10.1002/2014PA002772

Figure 7. Stratigraphic distribution of dinocyst assemblages, ecology indices (dominance, diversity, and equitability), P/G, C/M, AOM/Phyt ratios, and TOC.
Abbreviations: P/G = peridinioid to gonyaulacoid dinoagellate cysts; C/M = continental/marine palynomorphs; AOM/Phyt = amorphous organic matter to phytoclasts; TOC = total organic carbon. Dashed line = average values.

5.3. Dinocyst Ecology: Diversity, Equitability, and Dominance


The values for dinocyst diversity and equitability in well GTP-24-SE generally increase upward in section (Figure 7);
the lowest values are H(S) = 0.28 and E = 0.19. The maximum dinocyst diversity and equitability values occur in the
upper part of the section, where average values of H(S) = 2.14 and E = 0.97 have been calculated.
The lowest diversity and equitability values usually occur in intervals dominated by the Subtilisphaera
community. High diversity and equitability values tend to coincide with high abundances of the
Oligosphaeridium and Spiniferites communities (Figure 7). Dominance is inversely proportional to diversity
and equitability, including minima (D = 19.1%) in the middle to upper part of the section. In general, high
dominance values coincide with high abundances of the Subtilisphaera community (highest D = 93.5%).
High values also occur where another community dominates the sample, whereas low values are correlated
with a high abundance of the Oligosphaeridium community. The lowest dominance value coincides with the
highest value for the Oligosphaeridium community (37.5% of total communities at 65.9 m).
5.4. Peridinioid/Gonyaulacoid Index
In the studied section, only species of Subtilisphaera and Palaeoperidinium cretaceum are peridinioids. With
the exception of the ceratioid cysts Odontochitina and Pseudoceratium, all other taxa are gonyaulacoids.
Owing to the high abundance of Subtilisphaera senegalensis, the peridinioid/gonyaulacoid (P/G) ratio is
controlled by this single species. The P/G values are plotted next to the values for diversity, equitability, dominance, and communities (Figure 7). Maximum P/G values (or dominance of peridinioids) occur in the lower
part of the section, coinciding with the highest values of the Subtilisphaera community. Gonyaulacoids
display strong uctuations and become abundant from ~160 m upward.
5.5. Continental/Marine Index
Changes in the co-occurrence of terrestrial palynomorphs (pollen and spores) and those of marine origin
(dinocysts) in marine deposits can be used as an indicator of proximal-distal trends [Pellaton and Gorin,

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

13

Paleoceanography

10.1002/2014PA002772

2005], thus providing a means for


interpreting sea level changes and
their impact on marine environments
[Scott et al., 1984; Mudie, 1989;
Versteegh and Zonneveld, 1994;
Bombardiere and Gorin, 1998, 2000;
McCarthy and Mudie, 1998; Carvalho
et al., 2013]. In general, the ratio of
terrestrial to marine palynomorphs
decreases from nearshore to
offshore environments.
Terrestrial palynomorphs are present
in all samples of the section. Usually,
the terrestrial palynomorph curve
coincides with the Subtilisphaera
community and, consequently, with that of peridinioid cysts. Thus, the C/M ratio is highest in the lower part
of the section (Figure 7). In the middle and upper parts of the section, this curve shows a decreasing trend,
although with strong uctuations.

Figure 8. Canonical correspondence analysis (CCA) plot showing the ordination


direction of dinocyst communities and ecological parameters (equitability,
AOM/Phyt, productivity and terrigenous input).

5.6. Amorphous Organic Matter/Phytoclasts


Carvalho [2001] suggested that the AOM recorded in the GTP-24-SE section is primarily of marine origin,
on the basis of moderate to high uorescence, which is indicative of an aquatic origin and low oxygenation, in association with high abundance of marine palynomorphs. Thus, it is reasonable to assume
that high values of the ratio indicate marine environments with low oxygenation, whereas low values
reect a higher degree of terrestrial input. The AOM/Phyto ratio, similarly to the C/M ratio, is slightly
dominant in the lower part of the section (Figure 7). Above 110 m, the ratio shows an increasing
trend with particularly high values between 75 and 55 m, corresponding to high abundances of the
Cyclonephelium-Exochosphaeridium community. However, values of the C/M ratio decrease in the upper
part of the section.
The ratio of fern spores to xerophytic palynomorphs (Fs/X) was calculated by dividing the number of all in situ
(non-reworked) fern spores by the number of xerophytic forms (Classopollis + Equisetosporites + Steevesipollenites
+ Gnetaceopollenites) (Figure 8). High values of Fs/X reect humid conditions, as ferns depend on water
to reproduce.

6. Discussion
The upper Aptian succession of the Sergipe Basin is characterized by a transgressional trend, reecting changes
in environmental conditions over time. These changes are indicated by the lithologic characteristics of the
column and coincide with interpretations based on vertical distribution patterns of the Oligosphaeridium,
Cyclonephelium-Exochosphaeridium, Spiniferites, and Subtilisphaera dinocyst communities. The sharp break
between the Oligosphaeridium and the Cyclonephelium-Exochosphaeridium communities shown in the
dendrograms (Figure 5) reects the separation of open-marine (e.g., Oligosphaeridium spp.) and nearshore or
environmentally stressed dinocysts (e.g., Subtilisphaera spp., Cyclonephelium spp.).
Three of the communities, the Cyclonephelium-Exochosphaeridium, Spiniferites, and Subtilisphaera communities, show stratigraphically characteristic occurrences (Figure 7). The Oligosphaeridium community is
subordinate, with a very low abundance (1.5% of the total communities).
6.1. Subtilisphaera Community
In the lower (268180 m) and middle (11272.25 m) parts of the section, the prominent feature is the
dominance of the Subtilisphaera community, in particular of the species Subtilisphaera senegalensis.
This community makes up an average of 79.2% of all communities in the lower part of the section,
reaching 93.4% (183 cysts in 200 counted) at 190.75 m. In the middle part of the section the abundance
is 41.8% (Figure 7).

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

14

Paleoceanography

10.1002/2014PA002772

Table 3. Types of Subiltisphaera Ecozone Proposed by Arai et al. [1994]


Ecozones

Characteristics

Subtilisphaera Ecozone (S. senegalensis)


Subtilisphaera spp. Ecozone, with Spiniferites seghiris
Subtilisphaera cheit Ecozone
Subtilisphaera spp. Ecozone, diluted by other dinoagellates
Subtilisphaera spp. Ecozone, diluted by terrestrial palynomorphs

>70% of Subtilisphaera species


Subtilisphaera species associated with Spiniferites seghiris
Predominance of Subtilisphaera cheit
High abundance of Subtilisphaera and other dinoagellates
High abundance of Subtilisphaera and terrestrial palynomorphs

Arai et al. [1994, 2000] proposed Subtilisphaera ecozones (Table 3) for the Early Cretaceous of the incipient
South Atlantic. These ecozones were originally identied in the Aptian of the Cear Basin [Regali, 1989b]
and later in other continental margin basins of Brazil, including the Sergipe Basin [Carvalho, 2001, 2004].
According to Arai et al. [1994, 2000] the Subtilisphaera Ecozone is characterized by the predominance of
Subtilisphaera cysts, forming a nearly monospecic community. On the basis of the very high abundance
of Subtilisphaera, it can be concluded that the Subtilisphaera (S. senegalensis) Ecozone (Table 3) is present
in this interval.
Species of Subtilisphaera have been associated with restricted, low-salinity, marine environments [Jain and
Millepied, 1975; van Helmond et al., 2014]. These species are common in low-diversity communities [Arai,
2001; Arai et al., 1994, 2000; Lana, 1997; Lana and Pedro, 2000; Antonioli, 2001; Antonioli and Arai, 2002],
reecting an extensive bloom in mid-Cretaceous epicontinental seas. The distribution of Subtilisphaera
strongly suggests connection with the Tethyan Realm [Arai, 2007, 2009, 2014; Arai et al., 2000], although there
are reports of the genus also from the Southern Hemisphere [Morgan, 1980; Cookson and Eisenack, 1982].
The two intervals of the succession dominated by the Subtilisphaera community are characterized by low carbonate contents and a predominance of shales, suggesting that during these intervals a high source-area
relief supplied large volumes of siliciclastic sediments [Koutsoukos et al., 1993]. This condition also explains
the high content of terrestrial material, indicated by the high C/M ratios in this phase. Results of a canonical
correspondence analysis (CCA) also suggest that the distribution of the Subtilisphaera community is mainly
related to terrigenous input (high C/M; Figure 8). These results suggest a preferential distribution of this community in low-salinity conditions.
The paleoceanographic model proposed by Pross [2001] and Pross and Schmiedl [2002] to interpret paleooxygenation based on the species Thalassiphora pelagica presents conditions similar to the intervals in the
Sergipe succession, where Subtilisphaera senegalensis dominates. According to Pross [2001], horizons with
high abundances of Thalassiphora pelagica are characterized by low diversity and equitability, high continental supply, high productivity, low salinity, a nearly monospecic occurrence, and inverse proportions of
dinoagellates characteristic of open-marine environments (e.g., Spiniferites spp.). These factors are supported by low values of diversity (H(S) = 0.3) and equitability (E = 0.2), very high values of P/G and C/M
(Figure 7), high abundance of the Subtilisphaera community (79.2% of all communities), very low abundance
of the Spiniferites community and absence of the Oligosphaeridium community (Figure 7).
The high abundance of the Subtilisphaera community probably resulted from a more humid period,
increased freshwater input, a high rate of nutrient supply, and the formation of a pycnocline separating
slightly less saline surface waters from higher saline deeper waters. These conditions would have impeded
vertical circulation, as suggested by Pross and Schmiedl [2002] for a dominance interval of Thalassiphora
pelagica. Moreover, moderate to high AOM/Phyto ratios and moderate values of total organic carbon are
recorded (TOC; Figure 7). A high AOM content is indicative of a more restricted environment far from terrigenous sources, unlike the environment indicated, where high abundance of Subtilisphaera community is
recorded (Figure 7). Apparently, and despite the conditions of a restricted sea that allowed the accumulation
of AOM, the Tethyan inuence is conrmed by the presence of the Subtilisphaera Ecozone. In the following
interval of the section studied, where the Subtilisphaera community is dominant (11272.25 m), the relationship with the ecozone proposed by Arai et al. [1994] is also conrmed, including their Subtilisphaera spp.
Ecozone, diluted by other dinoagellates (Table 3). In this case, the Subtilisphaera community is associated
predominantly with the Cyclonephelium-Exochosphaeridium community and shows moderate to high
diversity indices. This association was also reported by Marshall and Batten [1988], supporting the view of
a shallow-marine environment with episodes of terrestrial input.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

15

Paleoceanography

10.1002/2014PA002772

6.2. Spiniferites Community


The Spiniferites community occurs in the middle (180112 m) and uppermost (37.516 m) parts of the section,
being more prominent in the middle interval (Figure 7), where it represents the rst signicant transgression
recorded in the section (Figure 3). The community probably reects the rst important ooding surface
observed in the section. In this part of the interval, there is a predominance of Spiniferites chebca.
The high values of Spiniferites spp. coincide with the carbonate-dominated lithofacies 2b (e.g., calcareous
mudstones). Koutsoukos et al. [1991] associated the deposition of calcareous mudstones with restricted shelf
environments (pelmicrites) and offshore, open-shelf to upper-slope conditions (biomicrites) during peak carbonate deposition at maximum relative sea level rises.
In the transition from the Subtilisphaera to the Spiniferites community, all measured parameters indicate an
increase in water oxygenation, particularly in the productivity P/G (average 0.3) and AOM/Phyto ratios,
whereas the terrestrial input (C/M) also decreases, though not signicantly. This indicates a relative rise in
sea level onto the shelf in this interval. However, the continuous input of terrestrial material in the marine
environment created conditions for the formation of the Subtilisphaera community, as evidenced by three
abundance peaks, at 167.3, 134.4, and 120.6 m. Moreover, these results show that the relative sea level rise
probably occurred in pulses, with the changes causing replacement of the communities. This view is supported by the moderate values of the diversity and equitability indices and by the dominance values exceeding the general average (55.9%) in half of the samples (eight samples) from this interval.
Conditions where the Spiniferites community is abundant in the uppermost part of the section (37.516 m)
differ from those of the middle part (180112 m). In the upper part, the community coexists primarily with
the Cyclonephelium-Exochosphaeridium and Oligosphaeridium communities, with a predominance of
Spiniferites ramosus. Here ecological parameters, such as diversity and equitability, show high values. The
below-average values coincide with the predominance of the Subtilisphaera community, demonstrating that
there were still episodes of terrestrial input into the basin, although of less intensity. In comparison with the
other communities, the Spiniferites community tends to be more dominant in intervals characterized by low
abundances of the Subtilisphaera community. CCA analysis (Figure 8) conrms that Spiniferites spp. tend to
decrease in abundance in environments with high productivity and terrestrial input, where Subtilisphaera
spp. are predominant.
In summary, it appears that only Spiniferites spp. were able to adapt rapidly to the new conditions, i.e., a more
open-marine environment with normal salinity.
6.3. Cyclonephelium-Exochosphaeridium Community
The Cyclonephelium-Exochosphaeridium community is conned to the upper part of the section, with prominent abundance peaks in the interval 72.2542.75 m (Figure 7). As mentioned above, this community already
shows moderate to high values in the previous interval, where the Subtilisphaera community dominates.
However, with increasing abundance, there was an abrupt decrease in the Subtilisphaera community, probably caused by a decrease or change in terrestrial input and/or a slight rise in sea level.
The highest values of the Cyclonephelium-Exochosphaeridium community coincide with an interval of dark
shales, the second highest TOC average (2.1%) and a moderate to high abundance of AOM. This combination
is close to the suggested environmental characteristics of the Cyclonephelium and CyclonepheliumExochosphaeridium association [Li and Habib, 1996; Lamolda and Mao, 1999; Harris and Tocher, 2003;
Barroso-Barcenilla et al., 2011; Peyrot et al., 2011; van Helmond et al., 2014]. As mentioned above, the
Cyclonephelium-Exochosphaeridium community is interpreted as indicative of nearshore to bathyal conditions
and in some cases of restricted marine environments (see Table 1). Furthermore, according to Lamolda and
Mao [1999], Cyclonephelium typically occurs in organic-rich marls or black shales, indicating a stressed
environment, as recorded by Lana [1997, 1998] for upper Cenomanian shales of the Potiguar Basin of
Brazil, with low levels of oxygen throughout the water column. This evidence points to a dysoxic-anoxic event
for this interval.
On the basis of micropaleontological, geochemical, and sedimentological studies, Koutsoukos et al. [1991] suggested the occurrence of intermittent dysoxic-anoxic events on the shelf during deposition of the Riachuelo
Formation. They associated the oxygen depletion with processes that restricted the free circulation of bottom

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

16

Paleoceanography

10.1002/2014PA002772

Figure 9. (a) Stratigraphic column showing the Cyclonephelium-Exochosphaeridium community versus C/M, AOM/Phyto, TOC, and fern spores to xerophytes (Fs/X).
Results of Rock-Eval pyrolysis for samples with high (orange dots) to moderate (yellow dots) content of Cyclonephelium, plotted in hydrogen index/oxygen index
diagram distinguishing between two phases. Classication of kerogen according to Erbacher et al. [1996]. Abbreviations as in Figure 7. Dashed line = average values.

ocean waters including physiographic barriers in the earliest opening of the setentrional South Atlantic Ocean,
salinity-stratied water masses, increased epipelagic primary productivity, and high sea levels. However,
Koutsoukos et al. [1991] did not associate these events directly with an Oceanic Anoxic Event (OAE).
Carvalho et al. [2006b] interpreted this part of the section as a highstand systems tract in the context of
sequence stratigraphy. From the base of the sequence, marked by a ooding surface at 257 m (see
Figure 3b), a clear decrease upward in dinocyst abundance and an increase in phytoclasts are observed.
However, a maximum-ooding surface at 46.25 m recorded by Carvalho et al. [2006b] is located within the
Cyclonephelium-Exochosphaeridium community interval and reects the phase of deepest marine conditions,
with low oxygenation values in of the studied section. This condition caused an increase in the TOC and AOM
values (Figure 7).
The interval of the Cyclonephelium-Exochosphaeridium community contains three pronounced peaks of the
Oligosphaeridium community (Figure 7). This suggests that the oxygen-depleted condition was interrupted
periodically during intervals of short-term oxygenation and normal salinity on the shelf during the initial
stage of deposition of the Riachuelo Formation.
Marshall and Batten [1988] differentiated the black shales of the Cenomanian-Turonian anoxic event from
those of the Aptian and Albian, suggesting that the Albian in the North Atlantic was dominated by terrestrially derived plant detritus. Erbacher et al. [1996] designated the levels with black shales of marine origin
P-OAEs (productivity oceanic anoxic events) and those of terrestrial origin D-OAEs (detrital oceanic anoxic
events). Thus, P-OAEs are related to transgressive periods with type II kerogen, whereas D-OAEs are related to
stillstands or falling sea levels and type III kerogen [Erbacher et al., 1996; Herrle et al., 2003].
The hydrogen-index/oxygen-index diagram for the Cyclonephelium-Exochosphaeridium community interval
(Figure 9) shows that most samples are very close to the type III kerogen eld, suggesting degradation of
terrestrial organic matter. The stratigraphic distribution of the Cyclonephelium-Exochosphaeridium community and hydrogen-index/oxygen-index diagram (Figure 9) distinguishes two intervals. The rst interval
(11272 m) is where the Cyclonephelium-Exochosphaeridium community becomes subordinate to the
Subtilisphaera community, even with high values of terrestrial material (C/M and phytoclasts), here associated
with high productivity. The second interval (7236 m) occurs in association with the CyclonepheliumExochosphaeridium community, which shows high AOM and TOC.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

17

Paleoceanography

10.1002/2014PA002772

In the interval of high Cyclonephelium abundance and increased AOM/Phyto ratio, a conspicuous increase in
the Fs/X ratio is observed (Figure 9). Humid conditions play an important role in dysoxic-anoxic events,
especially those linked to D-OAEs, when runoff is enhanced [Erbacher et al., 1996].
According to Leckie et al. [2002], multiple discrete black shales of OAE 1b are recorded in Mexico, the North
Atlantic basin (western Tethys) and the Mediterranean region (eastern Tethys). This late Aptian event has
been linked to cooling and a eustatic sea level fall. Therefore, in the context of relative sea level fall, the
nearshore and restricted-marine environments may have been formed as a result of structural physiographic
highs and lows. The Cyclonephelium-Exochosphaeridium community typical of these environments
increases accompanied by an increase in AOM abundance, signicant TOC peaks and a decrease in C/M
(Figure 9). However, the restricted environments periodically became more oxygenated (peaks of the
Oligosphaeridium community), with a simultaneous decrease of the Cyclonephelium-Exochosphaeridium
community and the TOC and AOM values.
In summary, the Cyclonephelium-Exochosphaeridium community is possibly related to a dysoxic-anoxic event,
although this community is of the D-OAE type of Erbacher et al. [1996]. This model is in agreement with the
intermittent dysoxic-anoxic conditions proposed by Koutsoukos et al. [1991] for the Riachuelo Formation.

7. Conclusions
Evaluation of the dinocyst assemblages of well GTP-24-SE reveals a paleoceanographic history for the late
Aptian of the Sergipe Basin marked by a progressive regional transgression. The change from an onshore
to open-marine environment was controlled by minor sea level uctuations during deposition of the upper
parts of the Muribeca Formation. This change was dominated by a progressive sea level rise at the onset of
deposition of the Riachuelo Formation.
1. The dinoagellate assemblages recorded in the studied section contain 28 moderately well preserved
species dominated by Spiniferites chebca and Subtilisphaera senegalensis. Less abundant, but still common
species are Cyclonephelium vannophorum, Spiniferites ramosus, and Exochosphaeridium phragmites.
2. Results from cluster analysis indicate the existence of four communities with distinct paleoecological
and paleoceanographic preferences, viz., the Oligosphaeridium, Cyclonephelium-Exochosphaeridium,
Spiniferites, and Subtilisphaera communities.
3. The Subtilisphaera community may be related to the Subtilisphaera Ecozone proposed by Arai et al. [1994],
with the lower part of the section corresponding to the Subtilisphaera (S. senegalensis) Ecozone and the
middle part to the Subtilisphaera spp. Ecozone, diluted by other dinoagellates, suggesting a restricted
to inner-neritic environment with Tethyan inuence.
4. Statistical analyses (CCA and cluster analysis) show a strong, signicant positive link between the contents of
the Subtilisphaera community and terrestrial palynomorphs as well as the Cyclonephelium-Exochosphaeridium
community; however, there is a signicant negative correlation with the Spiniferites and Oligosphaeridium
communities.
5. The rst signicant transgression recorded in the middle interval may reect the onset of the global
Aptian marine transgression, and is dominated by the Spiniferites community. The dark shales
enriched with amorphous organic matter represent the deepest marine environments in the column.
They are probably related to a dysoxic-anoxic event and are dominated by the CyclonepheliumExochosphaeridium community. The trends in the TOC and the Spiniferites/Cyclonephelium, AOM/Phyto,
C/M, and Fs/X ratios indicate short-term oxygen depletion resulting from the topography of the area.
However, most samples are classied as kerogen type III, suggesting a detrital origin (D-OAEs) for the dark
shales layers.
6. The uppermost part represents the installation of a fully Tethyan, open-marine environment, evidenced
by the dominant Spiniferites community.

Appendix A
The succession studied yielded a rich dinocyst assemblage. A percentage count of the 29 species identied in
the 87 palynological slides of well GTP-24-SE is given in the Appendix (Table A1).

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

18

16.00
16.65
18.63
19.35
19.85
20.53
20.85
22.80
22.88
23.55
26.23
26.83
28.00
30.60
32.25
33.20
35.95
37.55
40.10
42.25
42.65
46.35
47.90
48.60
49.70
50.60
51.80
52.35
54.50
56.10
57.60
57.80
58.70
59.30
60.70
65.45
65.90
67.00
67.40
71.60
72.25
75.40
77.20
80.40
82.55
83.00
84.90
86.70
88.40
91.25
94.35
97.65
98.20

0.0
5.4
4.8
16.7
5.9
0.0
22.9
8.8
52.8
36.4
0.0
0.0
0.9
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.7
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.1
10.0
0.0
1.1
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

36.0
48.6
27.6
16.7
41.2
5.0
54.2
0.0
6.7
18.2
0.0
3.2
22.0
0.0
4.3
5.1
27.1
8.6
36.8
3.9
14.1
20.0
15.3
23.1
32.7
14.7
27.6
2.1
56.4
32.8
53.3
4.5
57.5
23.1
56.7
21.7
6.3
0.0
3.3
26.2
52.6
22.2
0.0
11.8
2.8
4.8
0.0
5.7
5.0
8.5
0.0
13.9
15.2

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.7
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
11.3
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
2.7
0.0
0.0
0.0
0.0
2.1
0.0
0.0
0.0
0.0
3.2
0.0
0.0
18.4
0.0
45.8
0.0
0.0
2.0
63.5
44.0
43.1
5.5
25.2
60.3
24.1
16.7
10.3
34.5
23.3
37.3
3.4
11.5
15.9
42.2
6.3
14.3
36.7
52.4
12.6
3.7
8.3
23.7
34.7
11.3
36.6
70.1
2.5
40.7
11.5
20.8
15.2

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.7
0.0
0.0
0.9
0.0
0.0
0.0
0.0
1.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.5
1.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
8.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
5.6
0.0
0.0
0.0
1.5
0.0
0.0
0.0
0.0
0.9
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
11.5
0.0
0.0
0.0
0.0
4.6
0.0
0.0
1.7
12.5
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0

2.0
16.2
0.0
0.0
5.9
0.0
0.0
0.0
0.0
0.0
4.0
0.0
1.8
4.8
0.0
0.0
2.1
0.0
0.0
1.0
0.6
0.0
2.8
0.0
0.0
0.0
0.0
11.5
0.0
2.5
0.0
4.5
6.9
0.0
1.2
5.6
12.5
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
6.3
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.8
0.0
0.0
0.0
0.0
0.0
0.0
6.3
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

4.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.8
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
6.3
0.0
0.0
0.0
0.0
0.0
0.0
1.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

DEPTH Apteodinium Circulodinium Cribroperidinium Cyclonephelium Dinopterigium Exochosphaeridium Florentina Odontochitina Oligosphaeridium Oligosphaeridium Oligosphaeridium Oligosphaeridium Oligosphaeridium
(m)
granulatum distinctum
edwardsii
vannophorum
cladoides
phragmites
mantelli
operculata
albertense
complex
pocolum
pulcherrimum
totum

Table A1. Dinocyst Species Distribution (Percentages) of Well GTP-24-SE

Paleoceanography

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

10.1002/2014PA002772

19

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

12.0
2.7
11.4
0.0
0.0
0.0
2.1
0.0
1.1
9.1
20.0
0.0
4.6
0.0
2.8
10.1
4.2
2.6
0.0
0.0
0.0
0.0
0.0
2.2
0.9
2.9
0.0
0.0
0.0
0.8
0.0
0.0
0.0
0.0
4.3
0.0
0.0
2.9
3.3
2.4
3.2
0.0
0.0
4.3
0.0
1.6
0.0
0.0
5.0
0.0
1.1
1.0
0.0

12.0
21.6
25.7
11.1
0.0
0.0
8.3
1.5
10.1
9.1
36.0
25.4
40.4
28.6
55.3
63.3
0.0
38.8
24.1
1.0
3.2
20.0
18.1
1.1
25.2
13.2
10.3
1.0
5.1
0.8
1.1
1.5
0.0
26.9
6.7
2.8
0.0
8.6
10.0
4.8
5.3
3.7
0.0
2.2
4.2
3.2
0.0
14.9
5.0
0.0
0.0
22.8
33.3

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

6.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.1
0.0
0.0
0.0
11.0
4.8
5.7
0.0
0.0
6.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

20.0
2.7
8.6
38.9
11.8
5.0
0.0
13.2
16.9
18.2
8.0
1.6
15.6
57.1
0.0
17.7
8.3
25.0
24.1
4.9
16.0
6.0
12.5
59.3
11.2
7.4
23.0
18.8
15.4
21.0
20.0
9.0
25.3
0.0
10.4
23.3
6.3
11.4
36.7
7.1
25.3
1.9
0.0
16.1
9.7
53.2
2.4
8.0
0.0
13.6
0.0
27.7
27.3

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.9
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.4
0.0
0.0
0.0
2.3
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

8.0
0.0
2.9
11.1
35.3
90.0
10.4
72.1
10.1
9.1
32.0
65.1
0.0
4.8
12.1
3.8
12.5
18.1
6.0
87.3
0.6
0.7
0.0
7.7
2.8
0.0
4.6
34.4
12.8
5.0
2.2
43.3
2.3
38.5
4.3
2.2
31.3
54.3
10.0
7.1
1.1
68.5
91.7
40.9
47.2
4.8
58.5
0.0
72.5
37.3
83.9
1.0
6.1

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.6
6.3
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
3.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
19.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.8
0.0
0.0
0.0
0.0
0.0
9.0
0.0
0.6
8.7
4.2
1.1
1.9
0.0
5.7
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
9.7
0.0
0.0
0.0
0.0
0.0
11.9
3.0

Palaeoperidinium Pervosphaeridium Pseudoceratium Spiniferites Spiniferites Spiniferites Spiniferites Spiniferites Spiniferites Subtilisphaera Subtilisphaera Subtilisphaera Subtilisphaera Systematophora Tanyosphaeridium Trichodinium
cretaceum
spp.
securigerum
ramosus ancorifer
bejuii
chebca
lenzi
seghris
cheit
perlucida?
pirnaensis? senegalensis
cretacea
spp.
castanea

Table A1. (continued)

Paleoceanography

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

10.1002/2014PA002772

20

CARVALHO ET AL.

99.20
101.70
103.15
108.10
109.40
110.60
112.70
119.50
120.60
122.85
124.40
124.67
128.10
130.55
134.45
135.75
146.10
157.15
160.35
167.25
170.43
178.56
182.06
189.8
190.75
202.37
205.35
214.75
227.10
240.65
256.10
259.05
268.05
268.50

6.5
0.0
0.0
4.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
7.5
0.0
0.0
0.0
0.0
0.0

0.0
10.8
25.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
4.8
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.5
5.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
18.9
2.8
2.0
1.4
4.8
0.0
0.0
0.9
0.0
0.0
0.0
2.8
0.5
0.8
15.6
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
5.7
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
2.7
0.0
0.9
0.0
2.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.7
0.0
0.0
0.0
0.0

0.0
13.5
0.0
0.0
27.8
23.8
5.4
6.9
0.0
1.4
7.0
19.3
0.0
1.6
0.0
10.2
1.2
0.0
0.0
5.8
16.0
11.1
1.2
0.0
0.0
0.0
0.0
13.2
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.2
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

DEPTH Apteodinium Circulodinium Cribroperidinium Cyclonephelium Dinopterigium Exochosphaeridium Florentina Odontochitina Oligosphaeridium Oligosphaeridium Oligosphaeridium Oligosphaeridium Oligosphaeridium
(m)
granulatum distinctum
edwardsii
vannophorum
cladoides
phragmites
mantelli
operculata
albertense
complex
pocolum
pulcherrimum
totum

Table A1. (continued)

Paleoceanography

LATE APTIAN SOUTH ATLANTIC, DINOCYST

10.1002/2014PA002772

21

CARVALHO ET AL.

0.0
5.4
0.0
0.0
0.0
0.0
0.0
3.4
0.0
0.0
18.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.5
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
2.7
11.1
18.0
0.0
0.0
0.0
0.0
0.0
0.0
4.0
1.8
5.6
0.0
0.0
6.1
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

12.9
16.2
19.4
6.0
26.4
0.0
67.6
20.7
0.9
4.2
8.0
17.4
22.2
7.9
0.0
10.2
11.9
4.8
15.9
14.4
46.9
51.1
12.3
8.7
0.0
5.6
9.5
5.7
0.0
10.8
4.4
0.0
14.8
87.0

0.0
0.0
0.0
0.0
0.0
0.0
2.7
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.5
0.0
0.0
0.0
0.0
0.0
0.0
0.6
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

6.5
18.9
8.3
0.0
6.9
0.0
10.8
48.3
1.8
23.6
45.0
41.3
19.4
12.2
3.3
7.5
53.6
61.9
73.0
4.3
22.8
8.9
0.0
4.3
6.1
22.2
23.8
0.0
10.0
16.2
15.6
16.7
0.0
13.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
22.2
0.0
0.0
0.0
0.0
0.0
0.0
72.6
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.5
0.0
0.0
12.5
5.9
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.5
0.0
0.0
0.0
0.0
0.0
7.5
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
2.4
0.0
0.0
2.8
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
2.5
0.0
0.0
0.0
0.0
0.0

74.2
13.5
33.3
70.0
1.4
66.7
2.7
20.7
92.7
70.8
6.0
0.9
13.9
0.5
93.4
0.0
1.2
33.3
9.5
74.8
0.6
2.2
82.7
87.0
92.9
44.4
57.1
75.5
70.0
70.3
80.0
70.8
6.7
0.0

0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.2
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
0.0
2.4
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
1.2
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0
0.0

0.0
0.0
0.0
0.0
36.1
0.0
8.1
0.0
0.0
0.0
10.0
19.3
36.1
76.7
2.5
50.3
31.0
0.0
0.0
0.7
11.7
24.4
0.0
0.0
0.0
0.0
4.8
0.0
0.0
0.0
0.0
0.0
0.0
0.0

Palaeoperidinium Pervosphaeridium Pseudoceratium Spiniferites Spiniferites Spiniferites Spiniferites Spiniferites Spiniferites Subtilisphaera Subtilisphaera Subtilisphaera Subtilisphaera Systematophora Tanyosphaeridium Trichodinium
cretaceum
spp.
securigerum ramosus ancorifer
bejuii
chebca
lenzi
seghris
cheit
perlucida?
pirnaensis? senegalensis
cretacea
spp.
castanea

Table A1. (continued)

Paleoceanography

LATE APTIAN SOUTH ATLANTIC, DINOCYST

10.1002/2014PA002772

22

Paleoceanography
Acknowledgments
We are grateful to Javier Helenes
Escamilla (CICESE, Ensenada, Mexico) and
an anonymous reviewer, who improved
this paper signicantly with suggestions
and critical comments. We also thank
Susanne Feist-Burkhardt (SFB Geological
Consulting and Services, Ober-Ramstadt,
Germany) for additional palynological
analyses of the GTP-24 well (unpublished
Petrobras report), which provide a more
precise age assessment for the section.
We thank especially Marta Claudia Viviers
(Petrobras, Rio de Janeiro, Brazil), who
kindly analyzed the calcareous
microfauna in several samples of the
studied section. We express our thanks to
Petrobras for giving M.A. Carvalho the
opportunity to study the material. This
study was funded mainly by the Brazilian
National Council for Scientic and
Technological Development (Conselho
Nacional de Desenvolvimento Cientco
e Tecnolgico (CNPq), grant no.
302064/2010-9), the Research Support
Foundation of Rio de Janeiro State
(Fundao Carlos Chagas Filho de
Amparo Pesquisa do Rio de Janeiro
(FAPERJ), grant no. E-26/103.028/2008),
the Brazilian Research Funding
Organization (Coordenao de
Aperfeioamento de Pessoal de Nvel
Superior (CAPES), grant no. BEX
11616/13-0), and the German Academic
Exchange Service (DAAD), grant no.
A/13/03339). The data used in this study
can be accessed at http://archiv.ub.uniheidelberg.de/volltextserver/1586/1/
marcelo.pdf.

CARVALHO ET AL.

10.1002/2014PA002772

References
Abdel-Kireem, M. R., E. Schrank, A. M. Samir, and M. I. A. Ibrahim (1996), Cretaceous palaeoecology, palaeogeography and palaeoclimatology
of the Northern Western Desert, Egypt, J. Afr. Earth Sci., 22, 93112.
Antonioli, L. (2001), Estudo palinocronoestratigrco da Formao Cod Cretceo Inferior do Nordeste Basileiro, PhD thesis, Universidade
Federal do Rio de Janeiro, Rio de Janeiro, Brazil.
Antonioli, L. and M. Arai (2002), O registro da Ecozona Subtilisphaera na Formao Cod (Cretceo Inferior da Bacia do Parnaba, Nordeste do
Brasil): Seu signicado paleogeogrco, paper presented at 6 Simpsio sobre o Cretceo do Brasil, So Pedro-SP, Brazil, pp. 2530.
Arai, M. (1999). A transgresso marinha mesocretcea: Sua implicao no paradigma da reconstituio paleogeogrca do Cretceo no
Brasil, paper presented at 5 Simpsio sobre o Cretceo do Brasil, Serra Negra-SP, Brazil, pp. 577582.
Arai, M. (2001), Palinologia de depsitos cretceos no Norte e Meio-Norte do Brasil: Histrico e estado-de-arte, in O Cretceo na Bacia de So
Lus-Graja 2001, edited by D. F. Rossetti, A. M. Ges, and W. Truckenbrodt, pp. 175189, MPEG Editorao, Belm, Brazil.
Arai, M. (2005), Biodiversit des dinoagells de la marge brsilienne de lAtlantique Central et de lAtlantique Sud: Outil traceur des
changes entre lAtlantique Nord et lAtlantique Sud au Crtac moyen et suprieur, MSc thesis, Universit Pierre et Marie Curie, Paris.
Arai, M. (2007), Sucesso das associaes de dinoagelados (Protista, Pyrrhophyta) ao longo das colunas estratigrcas do Cretceo das
bacias da Margem Continental Brasileira: Uma anlise sob o ponto de vista paleoceanogrco e paleobiogeogrco, PhD thesis,
Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil.
Arai, M. (2009), Paleogeograa do Atlntico Sul no Aptiano: Um novo modelo a partir de dados micropaleontolgicos recentes, Bull.
Geocienc. Petrobras, 17, 331351.
Arai, M. (2014), Aptian/Albian (Early Cretaceous) paleogeography of the South Atlantic: A paleontological perspective, Braz. J. Geol., 44,
339350, doi:10.5327/Z2317-4889201400020012.
Arai, M., C. C. Lana, and E. Pedro (1994), Ecozona Subtilisphaera. Registro eocretceo de um importante episdio ecolgico do Oceano
Atlntico primitivo, Acta Geol. Leopold., 17, 521538.
Arai, M., J. Botelho Neto, C. C. Lana, and E. Pedro (2000), Cretaceous dinoagellate provincialism in Brazilian marginal basins, Cretaceous Res.,
21, 351366.
Arnaud-Vanneau, A., J. M. Bernaus, M. C. Raddaddi, and H. Arnaud (2008), Registration of the rst global Lower Aptian transgression
(Orbitolina marl level) in the paleotropics: Role of tectonics, climate and eustasy, paper presented at 33rd International Geological
Congress, Oslo, Norway.
Azevedo, R. L. M. (2004), Paleoceanograa e a evoluo do Atlntico Sul no Albiano, Bull. Geocienc. Petrobras, 12, 231249.
Barroso-Barcenilla, F., A. Pascual, D. Peyrot, and J. Rodrguez-Lzaro (2011), Integrated biostratigraphy and chemostratigraphy of the
upper Cenomanian and lower Turonian succession in Puentedey, Iberian Trough, Spain, Proc. Geol. Assoc., 122, 6781, doi:10.1016/
j.pgeola.2010.11.002.
Bengtson, P., and E. A. M. Koutsoukos (1991), Ammonite and foraminiferal dating of the rst marine connection between the central and
South Atlantic, in Lundadagarna I Historisk Geologi Och Paleontologi, Lund Publ. in Geol., vol. 96, p. 5.
Bengtson, P., and E. A. M. Koutsoukos (1992), Ammonite and foraminiferal dating of the rst marine connection between the central and
South Atlantic, in 1er Colloque de Stratigraphie et de Palogographie des Bassins Sdimentaires Ouest Africains, 11e Colloque Africain de
Micropalontologie, Bull. Cent. Rech. Expl. Mm., vol. 13, p. 403.
Bengtson, P., and W. Souza-Lima (2000), Os fsseis da bacia de Sergipe-Alagoas: Os amonides, Phoenix, 16, 18.
Bengtson, P., E. A. M. Koutsoukos, M. V. Kakabadze, and M. H. Zucon (2007), Ammonite and foraminiferal evidence for opening of the South
Atlantic Ocean in the early/mid Aptian (Early Cretaceous), in 51st Palaeontological Association Annual Meeting, Programme with
Abstracts, p. 20. (Also in The Palaeontological Association Newsletter, 66, Abstracts, p. 25.)
Bombardiere, L., and G. E. Gorin (1998), Sedimentary organic matter in condensed sections from distal oxic environments: Examples from the
Mesozoic of SE France, Sedimentology, 45, 771788.
Bombardiere, L., and G. E. Gorin (2000), Stratigraphical and lateral distribution of sedimentary organic matter in Upper Jurassic carbonates of
SE France, Sediment. Geol., 132, 177203.
Borchert, H. (1977), On the formation of Lower Cretaceous potassium salts and tachhydrite in the Sergipe Basin (Brazil) with some remarks on
similar occurrences in West Africa (Gabon, Angola etc.), in Time- and Strata-bound Ore Deposits, edited by D. D. Klemm and H. J. Schneider,
pp. 94111, Springer, Berlin.
Brinkhuis, H. (1992), Late Eocene to Early Oligocene dinoagellate cysts from central and northeast Italy, Ph.D. thesis, University of Utrecht,
Ultrecht, Netherlands.
Brinkhuis, H. (1994), Late Eocene to early Oligocene dinoagellate cysts from the Priabonian type-area (northeast Italy); biostratigraphy and
paleoenvironmental interpretation, Palaeogeogr. Palaeoclimatol. Palaeoecol., 107, 121163.
Brinkhuis, H., and W. J. Zachariasse (1988), Dinoagellate cysts, sea-level changes and planktonic foraminifers across the CretaceousTertiary
boundary at El Haria, northwest Tunisia, Mar. Micropaleontol., 13, 153191.
Brinkhuis, H., J. P. Bujak, J. Smit, G. J. M. Vesteegh, and H. Visscher (1998), Dinoagellate-based sea surface temperature reconstructions
across the CretaceousTertiary boundary, Palaeogeogr. Palaeoclimatol. Palaeoecol., 141, 6783.
Burger, D. (1982), A basal Cretaceous dinoagellate suite from northeastern Australia, Palynology, 6, 161192.
Carvalho, M. A. (2001), Paleoenvironmental reconstruction based on palynology and palynofacies analyses of the AptianAlbian in the
Sergipe Basin, northeastern Brazil, PhD thesis, Universitt Heidelberg, Heidelberg, Germany.
Carvalho, M. A. (2004), Palynological assemblage from Aptian/Albian of the Sergipe Basin: Paleoenvironmental reconstruction, Rev. Bras.
Paleontol., 7, 159168.
Carvalho, M. A., J. G. Mendona Filho, and T. R. Menezes (2006a), Palynofacies and sequence stratigraphy of the AptianAlbian of the Sergipe
Basin, Brazil, Sediment. Geol., 192, 5774.
Carvalho, M. A., J. G. Mendona Filho, and T. R. Menezes (2006b), Paleoenvironmental reconstruction based on palynofacies analysis of the
AptianAlbian succession of the Sergipe Basin, northeastern Brazil, Mar. Micropaleontol., 59, 5681.
Carvalho, M. A., R. R. C. Ramos, M. B. Crud, L. Witovisk, A. W. A. Kellner, H. P. Silva, O. N. Grillo, D. Riff, and P. S. Romano (2013), Palynofacies as
indicators of paleoenvironmental changes in a Cretaceous succession from the Larsen Basin, James Ross Island, Antarctica, Sediment.
Geol., 295, 5366, doi:10.1016/j.sedgeo.2013.08.002.
Cookson, I. C., and A. Eisenack (1982), Mikrofossilien aus australischen mesozoischen und tertiren Sedimenten, Palaeontogr. Abt. B, 184,
2363.
Courtinat, B. (1993), The signicance of palynofacies uctuations in the Greenhorn Formation (CenomanianTuronian) of the Western
Interior Basin, USA, Mar. Micropaleontol., 21, 249257.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

23

Paleoceanography

10.1002/2014PA002772

Courtinat, B., and A. Schaaf (1990), Les kystes de dinoagells enregistrent les variations doxygnation dans la zone photique. Le cas du
CnomanienTuronien du bassin vocontien, C. R. Acad Sci. Paris, Sr II, 311, 699704.
Davey, R. J. (1978), Marine Cretaceous palynology of Site 361, DSDP Leg 40, off southwestern Africa, Initial Rep. Deep Sea Drill. Proj., 40,
883913.
Davey, R. J., and J. Rogers (1975), Palynomorph distribution in Recent offshore sediments along two traverses off South West Africa, Mar.
Geol., 18, 213225.
Downie, C., M. A. Hussain, and G. L. Williams (1971), Dinoagellate cysts and acritarch associations in the Paleogene of southeast England,
Geosci. Man, 3, 2935.
Duane, A. M. (1994), Preliminary palynological investigation of the Byers Group (Late JurassicEarly Cretaceous), Livingston Island, Antarctic
Peninsula, Rev. Palaeobot. Palynol., 84, 113120.
Duane, A. M. (1996), Palynology of the Byers Group (Late JurassicEarly Cretaceous) of Livingston and Snow islands, Antarctic Peninsula: Its
biostratigraphical and palaeoenvironmental signicance, Rev. Palaeobot. Palynol., 91, 241281.
Erbacher, J., J. Thurow, and R. Littke (1996), Evolution patterns of radiolaria and organic matter variations: A new approach to identify
sea-level changes in mid-Cretaceous pelagic environments, Geology, 24, 499502.
Erdtman, G. (1943), An Introduction to Pollen Analysis, Ronald Press, New York.
Erdtman, G. (1969), Handbook of Palynology, Munksgaard (Scandinavian Univ. Books), Copenhagen.
Eshet, Y., S. Moshkovitz, D. Habib, C. Benjamini, and M. Magaritz (1992), Calcareous nannofossil and dinoagellate stratigraphy across the
CretaceousTertiary boundary at Hor Harar, Israel, Mar. Geol., 18, 199228.
Faegri, K., and J. Iversen (1966), Textbook of Pollen Analysis, Munksgaard (Scandinavian Univ. Books), Copenhagen.
Gedl, E. (1999), Lower Cretaceous palynomorphs from the Skole Nappe (outer Carpathians, Poland), Geol. Carpathica, 50, 7590.
Goodman, D. K. (1979), Dinoagellate Communities from the Lower Eocene Nanjemoy Formation of Maryland, U.S.A, Palynology, 3,
169190.
Goodman, D. K. (1987), Dinoagellate cysts in ancient and modern sediments, in The Biology of Dinoagellates, Botanical Monogr., vol. 21,
edited by F. J. R. Taylor, pp. 649722, Blackwell Scientic Publ., Oxford.
Hammer, ., D. A. T. Harper, and P. D. Ryan (2001), PAST: Paleontological statistics software package for education and data analysis,
Palaeontol. Electron., 4(1), 9.
Harding, I. C. (1990), Palaeoperidinium cretaceum: A brackish-water peridiniinean dinoagellate from the Early Cretaceous, Palaeontology, 33,
3548.
Harker, S. D., W. A. S. Sarjeant, and W. G. E. Caldwell (1990), Late Cretaceous (Campanian) organic-walled microplankton from the interior
plains of Canada, Wyoming and Texas: Biostratigraphy, palaeontology and palaeoenvironmental interpretation, Palaeontogr. Abt. B, 219,
1243.
Harland, R. (1973), Dinoagellates cysts and acritarchs from the Bearpaw Formation (upper Campanian) of southern Alberta, Canada,
Palaeontology, 16, 665706.
Harland, R. (1983), Distribution maps of Recent dinoagellate cysts in bottom sediments from the North Atlantic Ocean and adjacent seas,
Palaeontology, 26, 321387.
Harris, A. J., and B. A. Tocher (2003), Palaeoenvironmental analysis of Late Cretaceous dinoagellate cyst assemblages using high-resolution
sample correlation from the Western Interior Basin, USA, Mar. Micropaleontol., 48, 127148.
Hedlund, R. W., and G. Norris (1986), Dinoagellates cyst assemblage from middle Albian strata of Marshall County, Oklahoma, U.S.A, Rev.
Palaeobot. Palynol., 46, 293309.
Helby, R., and A. McMinn (1992), A preliminary report of early Cretaceous dinocyst oras from Site 765, Argo Abyssal Plain, Northwest
Australia, in Proceedings of the Ocean Drilling Program - Scientic Results, vol. 123, edited by F. M. Gradstein and J. N. Ludden,
pp. 407420.
Herrle, J. O., J. Pross, O. Friedrich, and C. Hemleben (2003), Short-term environmental changes in the Cretaceous Tethyan Ocean:
Micropalaeontological evidence from the Early Albian Oceanic Anoxic Event 1b, Terra Nova, 15, 1419.
Jain, K. P., and P. Millepied (1975), Cretaceous microplankton from Senegal Basin, W. Africa: Pt. II: Systematics and Biostratigraphy,
Geophytology, 5, 209213.
Karner, G. D., S. S. Egan, and J. K. Weissel (1992), Modeling the tectonic development of the Tucano and Sergipe-Alagoas rift basins, Brazil, in
Geodynamics of Rifting, Volume III. Thematic Discussions, Tectonophysics, 215, 133160.
Koutsoukos, E. A. M. (1989), Mid- to Late Cretaceous microbiostratigraphy, palaeo-ecology and palaeogeography of the Sergipe Basin,
northeastern Brazil, PhD thesis, Council for National Academic Awards, Polytechnic South West, Plymouth, U. K.
Koutsoukos, E. A. M. (1992), Late Aptian to Maastrichtian foraminiferal biogeography and palaeoceanography of the Sergipe Basin, Brazil, in
Biogeographic Patterns in the Cretaceous Ocean, Palaeogeogr. Palaeoclimatol. Palaeoecol., 92B, 295324.
Koutsoukos, E. A. M., and P. Bengtson (1993), Towards an integrated biostratigraphy of the upper AptianMaastrichtian of the Sergipe Basin,
Brazil, Doc. Lab. Geol. Lyon, 125, 241262.
Koutsoukos, E. A. M., and P. Bengtson (2007), Evaluating the evidence on the opening of the Equatorial Atlantic Gateway and its global
impact, Geol. Soc. Am. Annu. Meet., Abstr. Prog., 39(6), 445.
Koutsoukos, E. A. M., M. R. Mello, N. C. Azambuja Filho, M. B. Hart, and J. R. Maxwell (1991), The upper AptianAlbian succession of the Sergipe
Basin, Brazil: An integrated paleoenvironmental assessment, Am. Assoc. Pet. Geol. Bull., 75, 479498.
Koutsoukos, E. A. M., N. Destro, N. C. Azambuja Filho, and A. R. Spadini (1993), Upper Aptianlower Coniacian carbonate sequences in the
Sergipe Basin, northeastern Brazil, Am. Assoc. Pet. Geol. Mem., 56, 127144.
Lam, K., and R. Porter (1977), The distribution of palynomorphs in the Jurassic rocks of the Brora outlier, northeast Scotland, J. Geol. Soc.
(London), 134, 4555.
Lamolda, M. A., and S. Mao (1999), The CenomanianTuronian boundary event and dinocyst record at Ganuza (northern Spain), Palaeogeogr.
Palaeoclimatol. Palaeoecol., 150, 6582.
Lana, C. C. (1997), Palinologia e estratigraa integrada da seo Cenomaniano mdioTuroniano inferior da poro centro-leste da Bacia
Potiguar, NE do Brasil, MSc thesis, Universidade Federal do Rio Grande do Sul, Porto Alegre, Brazil.
Lana, C. C. (1998), The palaeoenvironmental distribution of the Upper Cenornanian-Lower Turonian dinocyst assemblages of Potiguar basin,
North-eastern.Brazil. - NTNU Vitenskmus. Rapp. bot. Ser. 19981: 90.
Lana, C. C., and E. Pedro (2000), Subtilisphaera almadaensis, uma nova espcie de dinoagelado da Bacia de Almada, Brasil, Rev. Univ.
Guarulhos, Geoci., V (Nmero Especial), 8688.
Lebedeva, N. K. (2008), Biofacies analysis of Upper Cretaceous deposits in the Ust-Yenisei Region: Implications of palynomorphs, Stratigr.
Geol. Correl., Geol. Korrel., 16, 8197.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

24

Paleoceanography

10.1002/2014PA002772

Leckie, R. M., T. J. Bralower, and R. Cashman (2002), Oceanic anoxic events and plankton evolution: Biotic response to tectonic forcing during
the mid-Cretaceous, Paleoceanography, 17(3), 1041, doi:10.1029/2001PA000623.
Legendre, P., and L. Legendre (1998), Numerical Ecology, 2nd ed, 853 pp., Elsevier, Amsterdam.
Li, H., and D. Habib (1996), Dinoagellate stratigraphy and its response to sea-level change in CenomanianTuronian sections of the Western
Interior of the United States, Palaios, 11, 1530.
Liengjarern, M., L. Costa, and C. Downie (1980), Dinoagellate cysts from the Upper EoceneLower Oligocene of the Isle of Wight,
Paleontology, 23, 475499.
Lignum, J., I. Jarvis, and M. Pearce (2007), The dinoagellate cyst record of the CenomanianTuronian boundary (OAE 2): Data from a newly
cored black shale succession, Wunstorf, northern Germany, Geophys. Res. Abstr., 9, 03854.
Lister, J. K., and D. J. Batten (1988), Stratigraphic and palaeoenvironmental distribution of Early Cretaceous dinoagellate cysts in the
Hurlands Farm Borehole, West Sussex, England, Palaeontogr. Abt. B, 210, 989.
Mahmoud, M. S., and A. S. Deaf (2007), Cretaceous palynology (spores, pollen and dinoagellate cysts) of the Siqeifa 1-X Borehole, Northern
Egypt, Riv. Ital. Paleontol. Stratigr., 113, 203221.
Marshall, K. L., and D. J. Batten (1988), Dinoagellate cyst associations in CenomanianTuronian black shale sequences of northern Europe,
Rev. Palaeobot. Palynol., 54, 85103.
McCarthy, F. M. G., and P. J. Mudie (1998), Oceanic pollen transport and pollen: Dinocyst ratios as markers of Late Cenozoic sea-level change
and sediment transport, Palaeogeogr. Palaeoclimatol. Palaeoecol., 138, 187206.
Mendes, J. M. C. (1994), Anlise estratigrca da seo neo-aptiana/eocenomaniana (Fm. Riachuelo) na rea do Alto de Aracaju e adjacncias
Bacia de Sergipe/Alagoas. MSc thesis, Universidade Federal do Rio Grande do Sul, Porto Alegre, Rio Grande do Sul, Brazil.
Mendona Filho, J. G., T. R. Menezes, J. O. Mendona, A. D. Oliveira, M. A. Carvalho, A. J. SantAnna, and J. T. Souza (2010), Palinofcies, in
Paleontologia, vol. 2, edited by I. S. Carvalho, pp. 283317, Intercincia, Rio de Janeiro, Brazil.
Millioud, M. E. (1967), Palynological study of the type localities at Valangin and Hauterive, Rev. Palaeobot. Palynol., 5, 155167.
Morgan, R. (1980), Eustasy in the Australian early and middle Cretaceous, Geol. Surv. New South Wales, Dep. Min. Res., Bull., 27, 1106.
Moustafa, T. T., and G. A. Lashin (2012), AptianTuronian palynomorphs from El-Waha-1 Well, southwestern part of the Western Desert,
Egypt, J. Appl. Sci. Res., 8, 18701877.
Mudie, P. J. (1989), Palynology and dinocyst biostratigraphy of the Late Miocene to Pleistocene, Norwegian Sea: ODP Leg 104, Sites 642 to
644, in Proceedings of the Ocean Drilling Program - Scientic Results, edited by O. Eldholm, J. Thiede, and E. Taylor, pp. 587610, Ocean
Drilling Program, College Station, Tex.
Musacchio, E. A. (2000), Biostratigraphy and biogeography of Cretaceous charophytes from South America, Cretaceous Res., 21, 211220.
Ogg, J. G., and L. A. Hinnov (2012), Cretaceous, in The Geological Time Scale, pp. 793853, Elsevier BV, Amsterdam.
Ojeda, H. A. O. (1982), Structural framework, stratigraphy and evolution of Brazilian marginal basins, Am. Assoc. Pet. Geol. Bull., 66,
732749.
Ojeda, H. A. O., and A. M. Fugita (1976), Bacia Sergipe/Alagoas: Geologia regional e perspectivas petrolferas, paper presented at XXVIII
Congresso Brasileiro de Geologia, Porto Alegre-RS, Brazil, pp. 137158.
Omran, A. M., H. A. Soliman, and M. S. Mahmoud (1990), Early Cretaceous palynology of three boreholes from northern Western Desert
(Egypt), Rev. Palaeobot. Palynol., 66, 293312.
Pacton, M., G. E. Gorin, and C. Vasconcelos (2011), Amorphous organicmatter Experimental data on formation and the role of microbes, Rev.
Palaeobot. Palynol., 166, 253267.
Pearce, M. A., I. Jarvis, A. R. H. Swan, A. M. Murphy, B. A. Tocher, and M. Edmunds (2003), Integrating palynological and geochemical data in a
new approach to palaoecological studies: Upper Cretaceous of Banterwick Barn Chalk borehole, Berkshire, UK, Mar. Micropalaeontol., 47,
271306.
Pellaton, C., and G. E. Gorin (2005), The Miocene New Jersey passive margin as a model for the distribution of sedimentary organic matter in
siliciclastic deposits, J. Sediment. Res., 75, 10111027.
Peyrot, D., F. Barroso-Barcenilla, E. Barrn, and M. J. Comas-Rengifo (2011), Palaeoenvironmental analysis of CenomanianTuronian dinocyst
assemblages from the Castilian Platform (NorthernCentral Spain), Cretaceous Res., 32, 504526, doi:10.1016/j.cretres.2011.03.006.
Peyrot, D., F. Barroso-Barcenilla, and S. Feist-Burkhardt (2012), Palaeoenvironmental controls on late Cenomanianearly Turonian dinoagellate
cyst assemblages from Condemios (Central Spain), Rev. Palaeobot. Palynol., 180, 2540, doi:10.1016/j.revpalbo.2012.04.008.
Powell, A. J., H. Brinkhuis, and J. P. Bujak (1996), Upper Paleocenelower Eocene dinoagellate cyst sequence stratigraphy of southeast
England, in Correlation of the Early Paleogene in Northwest Europe, Geol. Soc. Spec. Publ. London, vol. 101, pp. 145183.
Prauss, M. (2001), Sea-level changes and organic-walled phytoplankton response in a Late Albian epicontinental setting, Lower Saxony
basin, NW Germany, Palaeogeogr. Palaeoclimatol. Palaeoecol., 174, 221249.
Pross, J. (2001), Paleo-oxygenation in Tertiary epeiric seas: Evidence from dinoagellate cysts, Palaeogeogr. Palaeoclimatol. Palaeoecol., 166,
369381.
Pross, J., and G. Schmiedl (2002), Early Oligocene dinoagellate cysts from the Upper Rhine Graben (SW Germany): Paleoenvironmental and
paleoclimatic implications, Mar. Micropaleontol., 45, 124.
Regali, M. P. S. (1989a), Primeiros registros da transgresso neoaptiana na margem equatorial brasileira, paper presented at 11. Congresso
Brasileiro de Paleontologia, Curitiba-PR, Brazil, pp. 75293.
Regali, M. P. S. (1989b), Complicatisaccus cearensis: uma palinozona do Eocretceo do Brasil, paper presented at 11. Congresso Brasileiro de
Paleontologia, Curitiba-PR, Brazil, pp. 235274.
Regali, M. S. P., and P. R. S. Santos (1999), Palinoestratigraa e geocronologia dos sedimentos albo-aptianos das bacias de Sergipe e de
Alagoas, paper presented at 5 Simpsio sobre o Cretceo do Brasil, Serra Negra-SP, Brazil, 411420.
Regali, M. S. P., N. Uesugui, and A. S. Santos (1974), Palinologia dos sedimentos meso-cenozoicos do Brasil II, Bol. Tc. Petrobras, 17, 177191.
Regali, M. S. P., N. Uesugui, and A. S. Santos (1975), Palinologia dos sedimentos meso-cenozoicos do Brasil II, Bol. Tc. Petrobras, 17, 263301.
Rochon, A., A. de Vernal, J.-L. Turon, J. Matthiessen, and M. J. Head (1999), Distribution of Recent dinoagellate cysts in surface sediments
from the North Atlantic Ocean and adjacent seas in relation to sea-surface parameters, Am. Assoc. Stratigr. Palynol. Contrib. Ser., 35,
1146.
Sarjeant, W. A. S., T. Lacalli, and G. Gaines (1987), The cysts and skeletal elements of dinoagellates: Speculations on the ecological causes for
their morphology and development, Micropaleontology, 33, 136.
Scott, D. B., P. J. Mudie, G. Vilks, and D. C. Younger (1984), Latest PleistoceneHolocene paleoceanographic trends on the continental margin
of eastern Canada: Foraminiferal, dinoagellate and pollen evidence, Mar. Micropaleontol., 9, 181218.
Scull, B. J., C. J. Felix, S. B. McCaleb, and W. G. Shaw (1966), The inter-discipline approach to paleoenvironmenal interpretations, Gulf Coast
Assoc. Geol. Soc., Trans., 16, 81117.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

25

Paleoceanography

10.1002/2014PA002772

Seeling, J. (1999), Palaeontology and biostratigraphy of the macroinvertebrate fauna of the CenomanianTuronian transition of the Sergipe
Basin, northeastern Brazil with systematic description of bivalves and echnoids, PhD thesis, Universitt Heidelberg, Heidelberg,
Germany.
Skupien, P., and Z. Vaek (2002), Lower Cretaceous ammonite and dinocyst biostratigraphy and paleoenvironment of the Silesian Basin
(outer Western Carpathians), Geol. Carpathica, 53, 179189.
Souza-Lima, W., E. J. Andrade, P. Bengtson, and P. C. Galm (2002), A bacia de Sergipe-Alagoas: evoluo geolgica, estratigraa e contedo
fssil The Sergipe-Alagoas Basin: Geological evolution, stratigraphy and fossil content, Phoenix, Edio Espec., 1, 134.
Tappan, H. (1980), The Paleobiology of Plant Protists, Freeman, San Francisco, Calif.
Tyson, R. V. (1989), Late Jurassic palynofacies trends, Piper and Kimmeridge Oay formations, UK onshore and offshore, in Northwest European
Micropalaeontology and Palynology, British Micropalaeontol. Soc. Ser., edited by D. J. Batten and M. C. Keen, pp. 135172, Ellis Horwood,
Chichester.
Tyson, R. V. (1993), Palynofacies analysis, in Applied Micropalaeontology, edited by D. J. Jenkins, pp. 153191, Kluwer Acad., Dordrecht.
Tyson, R. V. (1995), Sedimentary Organic Matter: Organic Facies and Palynofacies, Kluwer Acad., Dordrecht, Holland.
Uesugui, N. (1979), Palinologia; tcnicas de tratamento de amostras, Bol. Tc. Petrobras, 22, 229240.
Van Helmond, N. A. M., A. Sluijs, J. S. Sinninghe Damst, G.-J. Reichart, S. Voigt, J. Erbacher, J. Pross, and H. Brinkhuis (2014), Freshwater
discharge controlled deposition of CenomanianTuronian black shales on the NW European epicontinental shelf (Wunstorf, North
Germany), Clim. Past Discuss., 10, 37553786.
Vandenberghe, N., H. Brinkhuis, and E. Steurbout (2003), The Eocene/Oligocene boundary in the North Sea area: A sequence stratigraphic
approach, in From Greenhouse to Icehouse: The Marine EoceneOligocene Transition, edited by D. R. Prothero, L. C. Ivany, and E. A. Nesbitt,
pp. 419437, Columbia Univ. Press, New York.
Versteegh, G. J. M. (1994), Recognition of cyclic and non-cyclic environmental changes in the Mediterranean Pliocene; A palynological
approach, Mar. Micropaleontol., 23, 147183.
Versteegh, G. J. M., and K. A. F. Zonneveld (1994), Determination of (palaeo-)ecological preferences of dinoagellates by applying Detrended
and Canonical Correspondence analysis to Late Pliocene dinoagellate cyst assemblages of the south Italian Singa section, Rev. Palaeobot.
Palynol., 84, 181199.
Vozzhennikova, T. F. (1965), Introduction to the Study of Fossil Peridinian Algae, translated by K. Syers, edited by W. A. S. Sarjeant, Boston Spa,
Yorkshire, England.
Wall, D., B. Dale, G. P. Lohmann, and W. K. Smith (1977), The environmental and climatic distribution of dinoagellate cysts in modern marine
sediments from regions in the North and South Atlantic Oceans and adjacent seas, Mar. Micropaleontol., 2, 121200.
Williams, G. L. (1977), Dinocysts. Their classication, biostratigraphy and palaeoecology, Ocean. Micropalaeontol., 2, 12311325.
Wilpshaar, M., and H. Leereveld (1994), Palaeoenvironmental change in the Early Cretaceous Vocotian Basin (SE France) reected by
dinoagellate cysts, Rev. Palaeobot. Palynol., 84, 121128.

CARVALHO ET AL.

LATE APTIAN SOUTH ATLANTIC, DINOCYST

26

You might also like