You are on page 1of 18

CHEMICAL

GEOLOGY
rwcr_uLvNc
ISOTOPE GEOSCIENCE
ELSEVIER Chemical Geology 130 ( 1996) 271-288

Global and local controls influencing the deposition of the La


Luna Formation ( Cenomanian-Carnpanian) , western Venezuela
Julio Perez-Infante a71,Paul Farrimond a7*, Max Furrer b
a Newcastle Research Group in Fossil Fuels and Environmental Geochemistry, Drummond Building, University ofNewcastle, Newcastle
upon Tyne, NE1 7RlJ, UK
b Lagouen, S.A., Filial Petroleos de Venezuela, Apartado 889, Caracas 1010 A, Venezuela

Received 12 April 1995; accepted 30 January 1996

Abstract

Bulk and molecular geochemical, micropalaeontological, and carbon-isotopic data are used to address the different local
and global factors influencing the environment of sedimentation of the La Luna Formation (Cenomanian-Campanian,
approximate palaeolatitude IS’N) in a single section in western Venezuela. Based on the constructed chronostratigraphic
framework, oxygen-depleted bottom-water conditions and black-shale deposition started in western Venezuela well before
the widespread occurrence of organic-rich sediments in higher palaeolatitude regions such as the Tethys and the North
Atlantic near or at the Cenomanian-Turonian boundary. In the La Luna Formation, palaeoenvironmental conditions that
allowed the preservation of organic matter (mainly of marine origin), prevailed until Santonian times in a distal platform
facies with very low siliciclastic input. Changes in lithology appear to reflect the local response to eustatic sea-level
variations and the presence of a migrating upwelling belt affecting the bioproductivity of silica and carbonate. A marked
8’3COrg isotopic excursion is recognised in the middle part of the section, and is apparently unrelated to local palaeoenviron-
mental changes in bioproductivity and oxygen depletion. Biological marker dam show no variations in association with the
isotopic excursion, being mainly controlled by local fluctuations in organic-matter input and preservation.

1. Introduction Ocean Drilling Program, and their correlation with


many exposed sections on land show that organic-
The Mid-Late Cretaceous age represents one of carbon-rich sediments of this age are confined to
the periods of more extensive distribution of particular stratigraphic horizons (for a review see
organic-rich sediments in both deep- and shallow- Arthur et al., 1990), leading to the conclusion that,
marine environments, throughout the world. Strati- within relatively narrow time envelopes (‘Oceanic
graphic data from the Deep Sea Drilling Project and Anoxic Events’), global oceanic conditions allowed
the preservation of large amounts of organic carbon
(Schlanger and Jenkyns, 1976; Jenkyns, 1980).
Scholle and Arthur (1980) have shown that temporal
* Corresponding author.
’Present address: Venezuelan Petroleum Research Institute, variations in 613C values correlate with these events.
Intevep, apartado 76343, Caracas 1070 A, Venezuela. Positive excursions in 613C of both organic matter

0009.2541/96/$15.00 Copyright 0 1996 Elsevier Science B.V. All rights reserved.


PII SOOOS-2541(96)00019-8
272 J. Perez-Infunre et al./Chemical Geology 130 (19961271-288

and primary carbonates have been widely observed ate-rich facies towards the Andes (southeast),
in many North American and European sections at or Colombia (west), and to the east.
near the Cenomanian/Turonian boundary (e.g. The La Luna Formation has often been quoted as
Arthur et al., 1987; Bralower, 1988; Gale et al., an example of C/T black shale deposition (i.e.
1993; Jenkyns et al., 1994). Although this isotopic Schlanger et al., 1987; Kuhnt et al., 1990). It is
excursion is commonly related to the presence of considered that dysaerobic to anoxic conditions pre-
black shales, at many localities the signal (measured vailed, and that these were probably linked to a
on primary carbonate) is found within Cenomaniar- regional anoxic event and seasonal upwelling along
Turonian shallow-water limestones lacking black the coast of South America (Tribovillard et al., 1991;
shales and deposited under oxic conditions (Schlanger Martinez and Hemandez, 1992). Macellari and De
et al., 1987). Therefore, this isotopic excursion is Vries (1987) proposed that the La Luna’s anoxic
considered a global event that resulted from a change sediments in northwestern South America are dis-
in the carbon-isotopic composition of the hydro- tributed on a Trough Sub-province (Colombia and
sphere and atmosphere (Scholle and Arthur, 1980; southwestern Venezuela), affected by dynamic up-
Popp et al., 1989). This has been explained through welling, and a Platform Sub-province (northwestern
widespread accumulation of organic matter (enriched Venezuela), ‘reflecting a global anoxic event’. Turo-
in 12C) in oceanic sediments leaving the oceanic nian-Coniacian organic-rich sediments are also found
carbon pool enriched in 13C and causing subsequent in Deep Sea Drilling Project sites in the Venezuela
carbonate or organic carbon deposits to be also Basin (Hay, 19851, and James (1990) argues that it is
enriched in 13C. This interval of extensive black ‘unreal’ to imagine that upwelling would have af-
shale deposition has been called the fected such a large area of northern South America
Cenomanian/Turonian Oceanic Anoxic Event and the southern Caribbean.
(‘OAE-2’; Schlanger and Jenkyns, 1976; Schlanger In this paper we present bulk and molecular geo-
et al., 1987) or the Cenomanian/Turonian Boundary chemical, micropalaeontological, and carbon-isotopic
Event (CTBE). It has been suggested that the anoxic data to address the occurrence of different local and
event is related to a major global sea-level rise. global events during the sedimentation of the La
However, the highest sea-level peak in the Late Luna Formation. In addition to defining the Cenoma-
Cretaceous appears to occur some time later in the nian/Turonian Boundary Event in the La Luna,
middle Turonian (between 91.5 and 90.3 million stratigraphic variability represented by lithofacies
years ago; Haq et al., 19871, as is pointed out by changes, biomarker composition and micropalaeonto
Hancock (1993) based on the stratigraphical position logical assemblages were considered in detail at a
of hardgrounds and ammonoid zones of the Tethyan single but representative location, to investigate tem-
region. poral relationships between eustatic changes and lo-
In western Mediterranean and North Atlantic re- cal bioproductivity of silica, carbonate and organic
gions the development of black shales is often re- matter.
stricted to the CTBE, but in low-latitude shelf basins
along the African continental margin, deposition of
black shales took place from as early as middle 2. Experimental
Cenomanian to late Turonian (Einsele and Wied-
mann, 1982; Kuhnt et al., 19901, suggesting that 2.1. Geologic setting and sampling information
local mechanisms contributed to their origin. In this
study, we consider another low-latitude occurrence Samples for this study were collected from Maraca
of mid-Cretaceous black shales: the La Luna Forma- Ravine (20 km southwest of Machiques City; Fig. l),
tion of Venezuela (Cenomaniar-Campaniann; palaeo- in the Perija Foothills. The La Luna here lies within
latitude of approximately 15”N). In ‘he western area an area of rain forest and although well exposed
of the Maracaibo Basin, the La Luna Formation during the dry season (November to March), some
mainly comprises alternating beds of marls and segments of the section are covered by dense vegeta-
pelagic limestones, changing laterally to less carbon- tion or stone blocks, producing gaps in sampling.
J. Perez-Infante et d/Chemical Geology 130 (1996) 271-288 273

The section was measured from the base of the La atomic absorbtion spectrometry using a Varian AA-
Luna Fm. This contact is abrupt and clearly distinc- 300 spectrometer, following digestion for 12 h at
tive as a lithological change between the massive 110°C with hydrochloric and hydrofluoric acids in
and organic-lean limestone of the Maraca Formation closed Teflon liners placed in stainless-steel cases.
and the first occurrence of organic-rich, laminated Total carbon and organic carbon (after acid dissolu-
marls of the La Luna Formation. A set of 90 samples tion of carbonate minerals) were determined using a
were collected from an estimated 160 m vertical Leco CS-244 Carbon-Sulphur Analyser. Carbonate
interval. Micropalaeontological analyses were per- content was determined by the difference between
formed on thin sections of selected samples placing the two measurements.
emphasis on radiolaria and planktic and benthic
foraminifers. 2.3. Sediment extraction and fractionation

2.2. Bulk geochemistry Approximately 25 g of powdered rock were ex-


tracted in a soxhlet apparatus using cellulose thim-
Samples were crushed to powder in a rotary disc bles for 48 h with dichloromethane (DCM)/methanol
mill (Terna@) and aliquots analysed by Rock-Eva1 (230 ml/20 ml) as solvent. Activated copper was
pyrolysis, and for their total organic carbon (TOC) added to the extraction flask to remove any elemen-
and carbonate contents. Free bitumen (S l), kerogen tal suiphur. After extraction, excess solvent was
pyrolysis yields (S2) and maximum temperature of removed by rotary evaporation and the dried extracts
pyrolysis CT,,,) were determined using a Leco were weighed in order to calculate the total ex-
THA200 Thermolytic Hydrocarbon Analyser. The tractable organic matter (EOM). The extracts were
elements silicon and aluminium were determined by fractionated by column chromatography using pre-

!
\
I
i
‘,
lo”, /
!
\
0 15km \
I I /’
? ?Field section studied
i

i
9”4O’N
I;

72”H
I
Fig. I. Sampling site of the La Luna Formation in the western Maracaibo Basin.
274 J. Perez-Infuntr et al./Chemical Geology I30 (1996) 271-288

extracted silica and alumina. Aliphatic hydrocarbons obtained following the original method described by
were obtained by eluting with petroleum ether. An Craig ( 1957) and modified by Hollander (1989). The
aromatic hydrocarbon fraction was obtained by elu- C-isotope composition of the obtained pure carbon
tion with 1: 1 petroleum ether:DCM, and the final dioxide gas was measured on a triple collector VG
fraction (polar compounds) with 1: 1 DCM:methanol. Micromass 903 Mass Spectrometer (ETH, Zurich),
using NBS-22 Hydrocarbon Oil Standard (613C =
2.4. Analysis of aliphatic hydrocarbons -29.63%0), and the 613C values are reported as per
mil (o/00)relative to the PDB isotopic standard. Preci-
Capillary gas chromatography was performed us- sion of the isotopic analyses (+0.20%0) was calcu-
ing a Carlo Erba Mega Series 5160, fitted with an lated using values obtained from duplicate analyses
OV-1 fused silica capillary column (30 m X 0.32 of samples.
mm id.; 0.25 km film thickness) and an on-column
injector. Analyses were performed using hydrogen as
the carrier gas, with an oven temperature of 50°C for 3. Results and discussion
2 min then ramped up at 4”C/min to a final tempera-
ture of 3OO“C, which was held for 20 min. Data were 3.1. Organic-carbon isotopic variation
acquired and processed using a VG Multichrom lab-
oratory data system. Subsequent analyses by gas In this study, isotopic-composition measurements
chromatography-mass spectrometry (GC-MS) were made on inorganic carbon in the samples of the La
performed using a Hewlett Packard 5890 gas chro- Luna Formation, indicate diagenetic alteration, par-
matograph fitted with an HP-5 fused silica capillary ticularly by secondary carbonate precipitation in the
column (25 m X 0.2 mm i.d.; 0.11 p_m film thick- sulphate reduction zone, giving considerable scatter
ness> and linked to a Hewlett Packard mass selective and unreliable negative values. Similar results have
detector (electron energy 70 eV; filament current 220 been reported by Schlanger et al. (1987) for the
p,A; source temperature 220°C). Helium was used as Cenomanian-Turonian Bridge Creek Limestone at
carrier gas. The initial oven temperature was 5O”C, Rock Canyon, Colorado. These authors associated
ramped to 300°C at 6”C/min, and held isothermally the scatter in the isotopic results of biogenic carbon-
at this temperature for 20 min. Compounds were ate to cyclically varying amounts of organic matter
identified by comparing their mass spectra (or ion in the section which caused diagenetic alteration of
responses) and relative retention times with either the primary signal of the skeletal calcite. Therefore,
those of reference compounds or with literature data. only organic carbon isotopic measurements made on
For quantification of GC-MS data, a deuterated isolated kerogen residues are considered in this study
sterane standard, (20R)-5o, 14o, 17a(H)-[2,2,4,4- of the La Luna Formation. These S13Corg values
D,]-cholestane, was added to the aliphatic hydrocar display a marked positive excursion from between
bon fractions (molecular ion m/z 376 and a base - 27.6 to - 28.0%0 and - 25.0 to - 25.5%0 in the 55
peak of m/z 221). No correction was made for the to 92 m interval of the section (Fig. 2). Visual
differences in mass spectral response of various hy- kerogen analysis and biomarker composition (see
drocarbons relative to the internal standard. later) show no significant changes at the boundaries
of this interval, indicating that the isotopic excursion
2.5. Carbon isotopes cannot be directly correlated to a change in organic
matter composition. In sediments of similar age from
S13C ratios of organic matter were measured from Deep Sea Drilling Project sites and in several on-
carbon dioxide gas generated by the combustion of shore sections in high latitudes, a similar isotopic
decarbqnated samples. An aliquot (1 to 5 g> of bulk shift occurs, often accompanied by a lithological
powdered sample, was treated with 1 N hydrochlo- change, an increase in the amount of TOC (e.g.
ric acid solution for 24 to 36 h, filtered with precom- Schlanger et al., 1987; Farrimond et al., 1990) and in
busted glass fibre filters (Whatman GF/C) and dried. the hydrocarbon generating potential of kerogen (HI).
The carbon dioxide from the organic fraction was However, in the La Luna Formation the isotopic
J. Perez-lnfante et al./Chemical Geology 130 (1996) 271-288 275

excursion is not associated with any marked change remains contribute increasingly to the sediment
in lithology, TOC or HI, and it occurs within a (Hilbrecht et al., 1992). On the other hand, total
thicker interval with good hydrocarbon source rock organic matter accumulated in sediments may not
potential (Fig. 2), suggesting that dysoxic-anoxic only be controlled by the primary productivity of the
bottom waters were already locally established in the overlying water column but also by local redox
basin. conditions. In an extensive review of the literature
Fluctuations in TOC and carbonate contents within related to the ‘productivity versus preservation de-
the interval of the isotopic excursion (Fig. 2) indicate bate’, Tyson (1995) pointed out that “in detail there
that local palaeoenvironmental conditions changed. is rather poor spatial correlation between areas of
However, isotopic values show no apparent correla- high marine productivity and areas of organic-rich
tion with these changes. Bulk parameters such as sediment deposition ( . . . ) due to the combined influ-
TOC and carbonate contents are directly influenced ence of other key factors including water depth,
by intra-basin processes affecting productivity and bottom water oxygenation, sediment grain size, down
preservation of biogenic carbonate, opal and organic slope redeposition and dilution and auto dilution
matter (Hut, 1988; Morse and Mackenzie, 1990; effects.” The lateral and stratigraphic heterogeneity
Ricken and Eder, 1991). Local biogenic productivity of the La Luna Formation (e.g. Macellari and De
can vary according to the relative proximity of a Vries, 1987; Martinez and Hemandez, 1992) sug-
specific site to upwelling systems. Modem coastal gests the existence of temporal and area1 variations
upwelling areas have a ‘core region’ of very high in nutrient-rich upwelling currents during deposition.
productivity and high nutrient abundance that is However, primary productivity may have had only a
commonly lo-20 km wide (Hilbrecht et al., 1992). minor influence on the final TOC signal in the site
Within these zones, the plankton is strongly domi- considered in this study since this parameter does not
nated by siliceous organisms (Bremner, 1983; Suess change markedly in those intervals where an increase
and Thiede, 1983; Molina Cruz, 1984). With increas- in primary productivity can be expected (i.e marl-
ing distance from this region the carbonate-produc- chert alternations of Unit 2). We interpret that TOC
ing organisms become more abundant, and their in the La Luna Formation was mainly controlled by

Unit
-
.
5
-
.
.
.
4 .
??? ?
a
- ??
. ??

3 8
.
- . 8
.
2 . L=
.
-
.O
.
1
.
- 0 I I q

-28 -25 -24 0 5 10 15

%CaCQ %TOC Carbonate free H.I. %SiOr/O/.Alz~

Fig. 2. S’3C(org) = carbon isotopic composition of organic matter (%o PDB); % carbonates (as CaCO,); % TOC (expressed on a
carbonate-free basis); H.I. = hydrogen index (from pyrolysis Rock-Eval, mg HC/g TOC); silica/alumina data of the La Luna Formation at
Quebrada Maraca. Lithological units are those proposed in this work. (Only values for marls and shales are plotted.)
276 J. Perez-Infante et al./Chemical Geology 130 (1996) 271-288

local environmental factors related to the preserva- increase in radiolarian abundance, the absence of
tion/degradation of organic matter. In contrast, the Rotalipora cushmani and the occurrence of hetero-
similarity of the La Luna Formation organic-carbon helicid/hedbergellid-dominated planktonic
isotopic excursion at Maraca Ravine with other sec- foraminiferal assemblages. The position of this
tions of North America, Africa and Europe suggest boundary between the R. cushmani and W. ar-
that these isotopic records reflect variation in the chaeocretaceu biozones in the La Luna Formation
global carbon cycle which responds to widespread (55 m) is additionally supported by a significant
geochemical changes in the hydrosphere and atmo- change in microfossil assemblage, including the first
sphere around the Cenomanian-Turonian Boundary appearance of Whiteinella balh’ca and Heterohelix
(Arthur et al., 1985; Hayes et al., 1989; Popp et al., reussi. The latter species is considered a morphotype
1989). of H. globulosa that, according to Nederbragt (199 1),
The timing and cause of the Cenomanian/ records the Cenomanian/Turonian boundary fauna1
Turonian isotopic event and its relationship with turnover. However, the placing of the Cenomanian-
microfossil assemblages has been a subject of dis- Turonian boundary depends upon which fossil groups
cussion for many years. Bralower (1988; U.S. West- are applied. The datums considered here are based
em Interior Basin) and Hilbrecht and Hoefs (1986; mostly on appearances and disappearances of spe-
NW Germany) found differences in the dating of the cific foraminifera. The biostratigraphy of Gale et al.
isotopic excursion based on nannofossils, suggesting (1993) largely employs macrofossils, and a compari-
that it was diachronous even on an intra-basin scale. son of our C-isotope curve with theirs indicates that
However, more recently, Gale et al. (1993) have the Cenomanian-Turonian boundary in the La Luna
shown convincing evidence that the timing and struc- Formation may actually be somewhat higher in the
ture of carbon-isotope curves of expanded Cenoma- section, nearer the top of Unit 3 (cf. Fig. 3). Unfortu-
nian-Turonian boundary sections in England (East- nately, our sampling resolution over this part of the
boume, Sussex) and North America (Pueblo, Col- section does not allow us to identify the detailed
orado) are identical with reference to first appear- C-isotope curve characteristics which appear later-
ances and disappearances of macro- and mi- ally correlatable (Gale et al., 1993; Jenkyns et al.,
cropalaeontological markers. 1994). Future macro- and micropalaeontological
Foraminifera and macrofossil assemblages appear studies and more detailed isotope stratigraphy of the
to be strongly influenced by palaeoenvironmental section need to be performed to support or correct
factors, and it seems likely that the fauna1 changes the interpretation proposed here.
observed in the La Luna Formation were to some At the top of the section, 6’“C values again shift
extent controlled by the local development of envi- gradually towards heavier values (Fig. 2), possibly
ronmental conditions that allowed the accumulation corresponding to a third anoxic event of the Creta-
of organic matter before the CTBE. Gale et al. ceous (OAE-3; Coniacian-Santonian period;
(1993) have pointed out that not all taxa of the Jenkyns, 1980; Arthur et al., 1990). This isotopic
Cenomanian-Turonian interval have identical ranges shift at the top of the La Luna Formation is consis-
at Pueblo (Colorado) and in Europe. However, a tent with the C-isotope curves for the Coniacian of
detailed correlation of the C/T isotopic event to the English Chalk and the Coniacian/Santonian of
planktonic foraminiferal zonation has been proposed the Italian Scaglia in the Bottaccione Gorge, Gubbio
by Kuhnt et al. (1990) for sections (e.g. Tarfaya (Jenkyns et al., 1994).
coastal basin, Morocco) deposited in similar palaeo-
latitudes (15”N) and water depths (200-300 m) to 3.2. Lithostratigraphic variation and sea-level
the La Luna Formation. They correlated the onset of changes
the isotopic shift in the Tarfaya to the top of the R.
cushmani biozone by the presence of important In the La Luna Formation, two major scales of
‘events’ that are also clearly recognisable at the base lithological variation can be distinguished both in the
of the isotopic excursion in the La Luna section field and from geochemical data. Marl-limestone
studied here (Fig. 3). These events include a marked and marl-chert alternations are easily recognisable
J. Perez-Infanie et al./Chemical Geology 130 (1996) 271-288 271

- -
?trE Planktonic Foraminifera
in Unit foraminifera relative n.y.O.
ctic zonation abundance
increase *
- - ........
..........
...........
............
............
5
.............
............
.............
............
D.ccmcavata .............
5 ............
- ............ - 08
140 .............
............
.............
............
.............
............
............
.............
............
............
............
.............
120 f ............
.............
4 3 D. primtiim
4 ............
.............
............
.............
............
-----_
............
g-z-- .............
............
......................

.............
............
- 89

100
3

I
............
H. helvetica
......................................
............
............. - 91
............
............
- ...........
...........
.........
.........
........
........
...... i
80 3
W. atchaeo-
cretacea 2
- - 92

60
2 - 92.:

40

R. cushmani

20

Fig. 3. Chronostratigraphic framework and biostratigraphic correlation used. ’Time scale of Haq et al. (1987). ’ Defined by first
appearance (F.A.) of Whiteinella baltica along with Heterohelix reussi (Ncderbragt, 1991); strong increase of radiolarian abundance,
positive 613C (organic) excursion, assemblages dominated by heterohelicid/hedbergellid, absence of Rotalipora cushmani (Kuhnt et al.,
1990); and decrease of planktonic foraminifera abundance around the C/T boundary (Kauffman, 1984). 3 Lower boundary defined by F.A.
of Marginotruncana coronata. 4 Lower boundary defined by F.A. of Dicarinella primitiva. ’ Lower boundary defined by F.A. of
Dicarinella concavata.

on a bed to bed scale (commonly < 1 m). Consider- been subdivided into five informal units in the pre-
ing the average sedimentation rates suggested for the sent study, based upon lithological changes (Table 1,
La Luna Formation (Martinez and Hemandez, 1992; Figs. 4 and 5):
and the present study; Fig. 4), the duration of such Unit I (the lowermost unit) is the most calcare-
bed-scale couplets in the La Luna Formation corre- ous, consisting of limestone layers (0.3 to 1 m thick)
sponds to 10 to 50 Ka, apparently within the Mi- interbedded with approximately 30% of thinner lami-
lankovitch frequency band. The bed-scale variations nated marls containing ellipsoidal carbonate concre-
are superimposed onto records of longer-term pro- tions. As in the rest of the section studied, the
cesses which affected the basin; it is these longer-term carbonate fraction is composed mainly of planktonic
variations which are addressed in this paper. foraminifera and calcareous nannoplankton.
The La Luna Formation at Quebrada Maraca has Unit 2 consists mainly of alternating layers of
278 J. Perez-lnfante et al./ Chemicd Geology 130 (1996) 271-288

T
etres Sedimentation t3.M. Accumulatiof 1
in Unit
rate (cm/l OOOy)2 rate (g/m2.y)3
dion

5 n.d. n.d
--_-
140

120-
4 1.5 (1.3) 1.1

lOO-

ao- 3 1.7 (1.5) 1.3

60-
2 5.5 (5.2) 5.1

40-

1 n.d n.d
20 -

m Shale pzJ Marlstone


lzz i%z”.a El Chetimarl
alternation

Fig. 4. Sedimentation and organic matter accumulation rates for the units of the La Luna Formation at Quebrada Maraca. Time zonation is
proposed on the basis of planktonic forammifera assemblages, scdimentological features and carbon isotope stratigraphy (see Fig. 3). ’ The
Haq et al. (1987) time scale has been used to calculate sedimentation rates. Differences of average compaction between units were
considered negligible. 3 The following formula has been used: O.M. AC = 0.1 X TOC X SR X D. D = 2.3 g/cm3 (average dry density
estimated for fine-grained, organic-rich rocks; Kuhnt et al., 1990).

well-laminated marls bearing abundant radiolarian are found in this unit, probably formed by storm
tests and thin bands of chert (up to 15 cm; approxi- action.
mately 15% of the unit). Unit 5 displays a gradual decrease in carbonate
Unit 3 lacks chert bands and the Si/Al ratio content, and a few thin chert bands ( < 2 cm) are
drops to a minimum (Fig. 2). This unit also has the observed.
lowest carbonate contents and the highest occurrence These changes in lithology appear to reflect the
of shale beds in the section (approx. 15% of this local response to eustatic sea-level variations and
unit). changes in the productivity of biogenic silica, proba-
Unit 4 sees an increase in carbonate content with bly controlled by variations in the relative proximity
an increase in the number of limestone beds. Rare of an upwelling belt to the site studied. Whilst local
thin arenaceous layers with abundant bivalve shells subsidence (in combination with sediment accumula-
J. Perez-lnfanre et al./ Chemical Geology 130 (1996) 271-288 219

Table 1
Average values (mean f standard deviation) of bulk geochemical parameters for the units of the La Luna Formation differentiated in this
study
Unit Metres n % CaCO, % TOC %TOC * H.I.

1 O-42 18 78.4 + 12.3 2.84 + 1.31 13.96 f 4.17 438 F 6 191 + 61


2 42-70 25 50.3 + 24.5 3.98 f 1.84 9.59 * 4.40 44Ok-3 214 rf: 62
3 70-91 14 31.3 5 23.0 3.29 + 1.29 5.94 + 3.99 439,l 210 + 33
4 91-140 19 70.1 + 17.2 3.02 f 1.53 10.89 + 4.32 439 f 2 274 + 83
5 140- 157 14 49.2 * 19.0 3.75 f 1.86 8.07 f 3.81 438 * 2 338 * 93

n = number of samples in each unit. Organic carbon content is additionally expressed on a carbonate-free basis (% TOC ’ ). HI. = hydrogen
index.

tion rate) could be an additional influence, Macellari America at this time. The middle Cenomanian-early
(1988) and Erikson and Pindell (1993) argue for Turonian comprises a transgressive period (Haq et
little regional tectonic activity in northern South al., 1987), followed by high-stand deposits corre-
sponding to the peak of maximum eustatic rise
(91.5-90.75 m.y. ago). Late Turonian to Santonian
Unit 5 times are characterised by a succession of transgres-
100
. sive and low-stand wedge systems tracts within a
80
long-term regressive period (Haq et al., 1987). The
following discussion examines geochemical and mi-
cropalaeontological records of sea-level changes in
the Maracaibo Basin, affecting the lithofacies pattern
of the La Luna Formation.
All the facies changes described above are consis-
tent with a pelagic belt in a distal pericratonic basin
as suggested for northwestern South America during
the Late Cretaceous by several authors (Zambrano et
80
Unit 3 al., 1971; Macellari, 1988; Martinez and Hemandez,
.
60 1992). Except for Unit 3, the La Luna Formation is
low in siliciclastic components (screening tests of
XRD analyses indicate clay content lower than 5%
in marl samples), supporting the hypothesis of very
Unit 2 distant sources of elastic sediment. Therefore, we
100
consider a low and fairly steady supply of detrital
80 II . .
input, with local changes in production and preserva-
tion of carbonate and opal, as being the most impor-
tant factors controlling sedimentation of the La Luna
Formation.
Unit 1
Carbonate sedimentation is strongly affected by
water depth, reaching its highest rates in clear, warm
and shallow waters. Consequently, the mass of bio-
genie carbonate grows most rapidly along preferred
positions on the upper part of any seaward slope,
% TOC within the zone of maximum biological productivity
Fig. 5. Plots of % TOC against % carbonates (as CaCO,) for the (see for example Wilson, 1975, and Morse and
units of the La Luna Formation at Quebrada Maraca. Mackenzie, 1990). At the site studied, the La Luna
280 J. Perez-infante et al./ Chemical Geology 130 (1996) 271-288

Fm. overlies platform deposits predominantly com- upwelling system in the area could have increased
posed of shallow-marine carbonates (Cogollo Group). both carbonate and biogenic silica productivity, the
The uppermost part of these shallow deposits (Maraca palaeolocation of the maximum carbonate and opal
Formation) consists of coarse bioclastic grainstones productivity belts need not have coincided since
with large, partly leached bivalves and echinoderm radio&a and calcareous planktonic organisms ap-
debris that were deposited during rapid deepening of pear to be highly tuned to specific oceanographic
a shallow platform environment (Vahrenkamp et al., environments (see Welling et al., 1992). Spumellar-
1993). Considering that elastic dilution by runoff ian and Nassellarian radiolarians, present in Unit 2,
was locally minimal (since positive areas were dis- suggest open-water conditions with a peak in abun-
tant) during the Late Cretaceous, we interpret the dance in middle-outer shelf or deeper upper-slope
units of La Luna with higher carbonate content environments (Koutsoukos and Hart, 1990).
(Units 1 and 4) to have been deposited on shallow In Unit 3, the absence of chert bands, preserved
outer-shelf/slope areas, while the carbonate-poor radiolarian tests and the lowest Si/Al ratios of the
Units (2 and 3) represent deeper environments and section (< 5; Fig. 2) shown for some beds of this
maximum sea-level periods. Although changes in unit, suggest that the inferred belt of high opal
nutrient supply can bring about profound changes in bioproductivity had shifted away before its deposi-
sedimentation rates independently of sea level tion. A striking decrease in the density of planktonic
(Gawthorpe et al., 1994), foraminifera and radiolar- foraminifera and the occurrence of dwarfed speci-
ian assemblages (see later), and the correlation of the mens of unkeeled foraminifera (Herdbergella and
units with global eustatic curves (Haq et al., 1987) Heterohelix) in Unit 3 (cf. Ford and Houbolt, 19631,
further support this interpretation. together with the fact that this unit shows the lowest
In the deeper-water units (2 and 3), foraminiferal average carbonate content of the section (31.3%;
shells are commonly well preserved, suggesting that Table l>, indicates that calcareous productivity also
the water depth never became sufficiently great to decreased. The linear sedimentation rates for the La
cause extensive dissolution of these organisms. Luna Formation units suggested in this study (Fig. 4)
(Berger (19701, based on suspended tests of must be considered as rough estimates since no
foraminifera from the East Pacific Rise, found slight correction has been made for compaction of the
dissolution of the foraminifera from 300 m.) In sediments. However, assuming similar average com-
general, the lithofacies of the La Luna Formation paction factors for the different units, the linear
remarkably resemble coeval pelagic sequences of the sedimentation rate appears to decrease by a factor of
North African Margin (e.g. Tarfaya coastal basin, 3 between Units 2 and 3.
Morocco; Einsele and Wiedmann, 1982) where max- Whilst a migrating upwelling system (moving in
imum palaeowater-depths of 300 m have been re- response to sea-level changes) could explain the
ported by Kuhnt et al. (1990). changes in productivity, an alternative explanation
As carbonate dissolution is unlikely in the studied could be the possible influence of an expanded oxy-
section, the lower carbonate contents of Units 2 and gen-minimum zone upon shallower parts of the wa-
5 suggest a relative decrease in primary carbonate ter column, affecting the development of intermedi-
productivity and/or dilution by a relative increase in ate planktic foraminifera (Martinez and Hernandez,
biogenic silica production. Opal bioproductivity is 1992) and radiolaria. Whatever its origin, the ob-
strongly controlled by the availability of nutrients, served decrease in planktonic productivity during the
and obviously, of dissolved silica in marine waters. accumulation of Unit 3 would decrease the sedimen-
In Unit 2, radiolarian tests and chert bands along tation rate, resulting in the deposition of sediments
with small-sized foraminifera (Het&ohelix, passively enriched with clays (Fig. 2); such sedi-
Globotruncana), high fauna1 density of Hedbergella ments are recognised as characteristics of
SP* and Globigerinoides and scarcity of Prae- maximum-flooding black shales (Wignall and May-
globotruncana, strongly suggest the presence of nu- nard, 1993).
trient-rich upwelling currents (Einsele and Wied- In Units 4 and 5, the facies patterns of Units 1
mann, 1982). Although the occurrence of a vigorous and 2 return, and may be associated with a rapid
J. Perez-lnfante et al./Chemical Geology 130 (1996) 271-288 281

regressive pulse during the middle Turonian (Haq et of inorganic material (biogenic and elastic). Al-
al., 19871, and the restoration of pelagic sedimenta- though average TOC values are similar for all units
tion on the outer platform. Normal-sized foraminifera of the La Luna Fm. (Table 11, plots of carbonate
are again observed, along with fragments of small content versus organic matter distinguish between
bivalves, particularly Znocerumus. The upper part of the units (Fig. 5). Unit 1 and (to a lesser extent) Unit
Unit 5 may also then represent the onset of a new 4 show inverse carbonate-organic carbon relation-
transgressive phase. The presence of an upwelling ships typical of deposition mainly dominated by
system (less vigourous than that of Unit 2) is sug- changes in the influx of carbonate components
gested by a similar foraminiferal assemblage and the (Ricken, 1994). Units 2 and 5 show more complex
occurrence of a few chert bands. patterns due to greater dilution by biogenic silica in
relatively deeper environments. Unit 3 shows a pat-
3.3. Organic matter enrichment in the La Luna tern largely independent of the carbonate concentra-
Formation tion, suggesting (as discussed earlier) that contribu-
tion of biogenic siliceous and calcareous components
Although the La Luna Fm. displays area1 and remained low during the deposition of this unit. It is
stratigraphic changes in lithology throughout the interesting to note that this apparent change in bulk
Maracaibo Basin, its organic matter content remains geochemistry, probably related to palaeoproductivity,
relatively high both in exposed sections at the edges occurs within the positive carbon-isotope excursion.
of the basin (Perija Foothills and northern flank of Despite variations in lithology, microscopic anal-
the Venezuelan Andes) and in cores from wells ysis of kerogens from the studied section show little
drilled all over the basin (Gonzalez de Juana et al., variation in the proportion of structured and amor-
1980; Talukdar et al., 1985). Contemporary organic- phous organic matter. All the samples examined are
rich facies are also found in eastern Colombia (La ,dominated by amorphous kerogen (95% or more>
Luna and Villeta Formations; Zumberge, 1984) and which has been widely ascribed to marine planktic
eastern Venezuela (Guayuta Group; Talukdar et al., and/or bacterial material (Tissot and Welte, 1984).
1985) suggesting regionally widespread anoxic bot- There is no palynological evidence for fluctuations
tom waters during the Late Cretaceous in northwest- in the relative proportions of marine and terrestrial
em South America. organic input during the sedimentation of the La
At Quebrada Maraca, the conformable contact Luna Formation.
between the La Luna Formation and the underlying,
shelly organic-lean limestones of the Maraca Forma-
tion, displays a striking increase in organic-matter 3.4. Biomarker composition
content, suggesting a rapid decrease in oxicity. Total
organic carbon (TOC) contents of the marl samples Signals preserved in the distributions of biomarker
vary from 1.50% to 6.85%, but the hydrocarbon-gen- compounds in a heterogeneous section such as the
erating potential (HI) shows less variation (Fig. 2). La Luna Formation are often sensitive to both envi-
Although maturation of organic matter greatly con- ronmental (source of organic matter, ecological as-
tributes to diminish original fluctuations in HI, the semblages, redox conditions, salinity) and diagenetic
data are consistent with relatively constant dysaero- factors (maturation, mineral matrix catalysis). In or-
bit to anoxic conditions throughout most of La der to minimise the influence of lithological varia-
Luna’s sedimentation. This interpretation is further tion on the organic matter in the section studied, only
supported by the near absence of benthic marl samples will be considered here. Furthermore,
foraminifera, with only a few occurrences towards organic matter maturity differences are negligible
the top of the formation. between the top and the bottom of the section;
Fluctuations in TOC can be produced by the neither T,,, values (440-445°C) nor vitrinite (very
combined effect of the flux of organic matter from scarce particles; reflectance 0.7-0.8% R,) display
the water column, its degree of preservation in sedi- any gradient. Therefore, variations in the biomarker
ments, and dilution by changes in the accumulation data are considered to be largely dependent upon
282 J. Perez-Infante et al./ Chemical Geology 130 (1996) 271-288

changes in the primary input of organic matter and Table 2


other palaeoenvironmental factors. Range of concentration (ppm of EOM) for selected compound
groups in the La Luna Formation
As noted earlier, and reported for other sections
UP Hopanes (C,,-C,,) 240-730
of the La Luna Formation in the northwest of the
aI3 Hopanes CC,, -C,J 145-725
Maracaibo Basin (Talukdar et al., 1985), molecular Total hopanes 385-1450
geochemistry and microscopic analyses indicate that Tricyclics (C ,9 -C,,) 180-315
the bulk of the organic matter is algal and bacterial Tricyclics (C,,-C,,) 330-1340
in origin. Saturated hydrocarbons (between 20% and Total tricyclics 50.5-1880
cxp@ Steranes CC,, -C,,) 120-360
30% of EOM) display little apparent molecular vari-
ation through the section. Normal and branched alka-
nes show a dominance of medium molecular weight
components (with a maximum around n-C,,) and
the pristane/phytane ratio is always lower than 0.8. throughout the section are tricyclic terpanes and
(Fig. 6). The most abundant cyclic compounds hopanes, whilst steranes are relatively low (Table 2).
The sterane distributions (C2,-C2s) display very lit-
tle variation, being always dominated by C,, ster-
anes, consistent with a marine source of organic
matter (Volkman, 1988), and suggesting a relatively
constant biotic input.
Principal component analysis (PCA) of biomarker
25 data from GC-MS analyses was used to aid interpre-

Ill
tation of molecular differences between samples.
This type of statistical analysis allows a complex
multivariate data set to be simplified by identifying
the covariance of individual variables (biomarkers)
and discriminating or grouping samples on the basis
of the direction and amplitude to which the distribu-
tions of the variables deviate from that of the aver-
age (see Davis, 1986). This method allows a large
data set to be expressed in terms of a limited number
of components (PCl, PC2, PC3, etc.) which are
19
linear combinations of the individual variables. In
this study, GC-MS peak areas of tricyclic terpanes,
hopanes and steranes (Table 3) were used for the
analysis, making a total of 42 variables for each of
the 27 samples selected. Biomarkers were only se-
lected if they could be accurately integrated for each
sample and suffered no co-elution problems. The
original raw data were normalised (to make each
sample the same ‘size’), autoscaled (i.e. dividing
each variable by its standard deviation, in order to
give them all comparable weight in the analysis) and
centred prior to PCA.
A ‘loadings plot’ can be used to show the rela-
~ Retention time -* tionships between variables (biomarkers) and the
degree to which each variable contributes to the PC’s
Fig. 6. Typical gas chromatograms of the La Luna Formation at
Quebrada Maraca. Pr = pristane; Ph = phytane; selected n-al- (Fig. 7). The first and second principal components
kanes are labelled with their carbon numbers. account for 70% of the total variance within the
J. Perez-lnfante et al./Chemical Geology 130 (1996) 271-288 283

scaled data set so they are the only ones discussed tricyclic terpane@Fs may be Tasmanites, a genus of
here. Long-chain tricyclic terpanes (> C,,) display extinct unicellular green algae, frequently associated
high positive loadings on PCl, identifying a strong with anoxic black shales (Tyson, 1995). Specifically,
importance of these compounds which vary largely extended tricyclic terpane hydrocarbons have been
independently from the other biomarkers considered identified as the major biomarkers in the bitumen of
(< C,, tricyclics, hopanes and steranes). In confir- the Tasmanian tasmanite (Simoneit et al., 1990).
mation, concentrations of long-chain tricyclic ter- However, remains of Tasmanites were not observed
panes (Fig. 8) closely follow the trend of PC1 varia- in the samples studied of the La Luna Formation.
tion through the section, with an important increase Furthermore, in the La Luna Formation the increase
in absolute abundance between 35 and 110 m of the in abundance of tricyclic terpanes ( > Cz6) is not
section. Tricyclic terpanes containing up to at least correlated to marked changes in any potentially re-
35 carbon atoms are clearly observed within Units 2, dox-sensitive parameters (i.e. HI, pristane/phytane).
3 and 4 (Fig. 9). Peters and Moldowan (1993) have Hopanes and steranes are important variables in
proposed that tricyclic terpanes may be source indi- PC2 (Fig. 7). The observed depth trend of PC2
cators relating to a group of bacterial or algal lipids, essentially records fluctuations in steranes vs.
but their specific origin is still poorly understood. hopanes, a parameter commonly associated with
Several authors (Volkman et al., 1989; Simoneit et changes in organic matter supply (e.g. Farrimond et
al., 1990; Aquino Neto et al., 1992; McCaffrey et al., al., 1994). With the exception of the sample at the
1994) have pointed out that at least one source of bottom of the section which displays a relatively

0.25
T PC2

C2s-C2s Tricyclics

023
24.26,27
b
8262936
e&t

035
C&II& Tricyclics 025
I

($1 -C23 Tricyclics

05

C&-C= Hopanes
013 /

Hopanes

Fig. 7. Loadings plot from the principal components analysis, showing the relationship between the biomarker variables in terms of the first
two principal components.
284 J. Perez-Infante et al./Chemicd Geology 130 (1996) 271-288

greater content of steranes, the sterane/hopane ratio not genetically associated with, the isotopic shift
displays a smooth trend throughout the section. This recorded in the organic matter.
is consistent with other geochemical results of this In general, none of the biomarker trends correlate
study (Rock-Eva1 pyrolysis, GC analyses and optical with the interval of the isotopic shift in the section.
microscopy) that show no evidence of changes in the Since the variation in biomarker composition is
proportion of terrestrial/marine organic matter sup- known to be largely controlled by local environmen-
ply during La Luna’s sedimentation. tal factors (e.g. palaeowater depth, productivity, local
The abundance of 28,30-bisnorhopane (BNH) also anoxia and salinity), the lack of parity with the
shows a clear trend through the section, similar to organic-C isotopic event is consistent with the latter
that shown by the long-chain tricyclic terpanes (Fig. being mainly independent of local conditions but
8). After a marked increase, the absolute concentra- rather a result of global depletion in ‘*C in the
tion of BNH falls from a maximum of 70 ppm oceans in response to widespread organic matter
(EOM) at the top of Unit 1 and the lower part of sequestration into sediments around the Cenoma-
Unit 2 to 10 ppm of the EOM in the upper part of nian/Turonian and Coniacian/Santonian bound-
Unit 4. The origin and significance of this compound aries.
remain unclear, although kerogen pyrolysis studies
indicate that it occurs in sediments as the free hydro-
carbon and is not present within the kerogen (Noble 4. Conclusions
et al., 1984). High abundances of BNH have been
associated with severely oxygen-deficient conditions Oxygen-depleted sedimentation of the La Luna
(Mello et al., 1989, and references within). In the La Formation in the western Maracaibo Basin started
Luna Formation maximum concentrations of BNH well before the CTBE, during the middle-late Ceno-
occur in Unit 2, where we interpret the highest manian in an open-marine environment, with a
productivity within the studied section, perhaps con- palaeowater depth of a few hundred metres and very
sistent with productivity-induced anoxia. Whatever low elastic input. We consider this early deposition
its origin, the presence of BNH in the La Luna of black shales to be the local response to a combina-
Formation clearly records changing local environ- tion of rapid sea-level rise and the presence of
mental conditions which precede, but are apparently upwelling currents. Geochemical evidence suggests

Tdcyclii retpanes’ PC1 Sleranesil-fopanes” 28.30 bisnorhopane


160

140

120

100

eo

so

40

20

i
n
“0 loo0 -10 0 10 0.15 0.35 0.55 10 0 10 0 50 100

pprn of EOM pprn of EOM

Fig. 8. Depth plots of the scores of the first two principal components identified from the biomarker data compared to selected molecular
profiles. The PC1 plot essentially records fluctuations in relative concentrations of long-chain tricyclic terpanes, while PC2 records
fluctuations in steranes vs. hopanes. The distinctive trend of absolute concentration of 28,30-bisnorhopane is also shown. ’ Css-Css.
* * Steranes/hopanes ratio = Speak areas aPP steranes (C,,-C,,)/Xpeak areas oP hopanes (C,,-C,,).
J. Perez-Infante et d/Chemical Geology 130 (1996) 271-288 285

that high carbonate and silica bioproductivity contin- at the site studied probably due to migration of
ued until the early Turonian. upwelling currents as is inferred from geochemical
The CTBE is, nevertheless, recorded in the mid- and micropalaeontological changes. However,
dle part of the La Luna Formation as a marked preservation of organic matter remained relatively
613C isotopic excursion. This isotopic shift is high due to the expansion of the oxygen-minimum
appg&tly unrelated to locally controlled changes in zone during the peak of the sea-level rise, as indi-
bioproductivity and intensity of oxygen depletion. cated from foraminiferal assemblages.
In the early Turonian, before the end of the Sedimentological and micropalaeontological evi-
isotopic excursion, biogenic productivity decreased dence suggest that the upper part of the La Luna

7
a

15

16

Fig. 9. Hopanes
Table 3.
JLL
and tricyclic terpanes (m/z 191) in typical samples from Units 2-3 (b) and Units 4-5 (a). Peak identities are given in
286 J. Perez-Infante et al./Chemical Geology 130 (1996) 271-288

Table 3 Formation (middle-Turonian to Santonian) experi-


Biological marker compounds used in the principal components
enced a rapid regressive pulse followed by a second
analysis (the numbers are used as peak labels in Figs. 7 and 9)
period of milder upwelling influence. Isotopic
Hopunes:
stratigraphy at the top of the La Luna Fm. indicates
I. 18a(H)-nisnomeohopane (Ts) the onset of a positive carbon isotopic excursion
2. 17o(H)-trisnorhopane (Tm) related to OAE-3 (Coniacian-Santonian).
3. 28,30-bisnorhopane
Biological marker data appear generally similar
4. C,, 17o (H), 21E (H) hopane
5. C,, 17o (H), 21B (H) hopane
for all samples upon initial inspection, although ap-
6. C,, 17~ (HI, 21p (H) hopane (22s) plication of principal components analysis identified
7. C,, 17o. (H), 21f3 (H) hopane (22R) some significant variability, particularly in the
8. C,, 17~ (H), 21p (HI hopane (22s) amount of extended tricyclic terpanes, 28,30-bi-
9. C,, 17a (H), 2lB (HI hopane (22R)
snorhopane and the sterane/hopane ratio. However,
IO. C,, 17a (H), 2lp (H) hopane (22s)
11. C,, 17a (H), 2lB (H) hopane (22R)
these biomarker variations do not occur in associa-
12. C,, 171~(H), 2lp (H) hopane (22s) tion with the isotopic excursion and are probably
13. C 34 I7a (H), 2 1p (HI hopane (22R) controlled by local fluctuations in organic matter
input and preservation. PCA identifies the abundance
Tricyclic terpanes:
of extended ( > C 27) tricyclic terpanes as the major
14. C ,9 13P(H),l4cx(H) tricyclic terpane source of biomarker variability. The significance of
15. C,, 13P(H), I4a(H) tricyclic terpane these compounds remains unclear, but they are ap-
16. C,, 13B(H),14a(H) tricyclic terpane
parently unrelated to the shorter-chain tricyclics, and
17. C,, 13P(H),14o(H) tricyclic terpane
18. Cz4 13@(H), 14a(H) tricyclic terpane
appear to have a different biological source.
19. C,, 13P(H),l4u(H) tricyclic terpane
20. C,, 13f3(H),l4cx(H) tricyclic terpane (22s)
21. C,, 13B(H),14a(H) tricyclic terpane (22R) Acknowledgements
22. C,, 13P(H),l4o(H) tricyclic terpane (22s)
23. C,, 13E(H),14u(H) tricyclic terpane (22R)
24. C,, 13E(H),14cx(H) tricyclic terpane (22% The authors are grateful to INTEVEP for financial
25. C,, 13B(H),l4a(H) tricyclic terpane (22R) support and a Ph.D. studentship (J.P.1). We also
26. C,, 13P(H), 141x(H) tricyclic terpane (22s) thank W. Scherer, M. Alberdi and A. Pilloud (IN-
27. C,, 13P(H),14a(H) tricyclic terpane (22R)
TEVEP) for their assistance during collection of
28. C,, 13P(H),14u(H) tricyclic terpane (22s)
29. C,, 13P(H),l4o(H) tricyclic terpane (22R)
outcrop samples. GC and GC-MS technical assis-
30. C,, 13E(H),14a(H) tricyclic terpane (22s) tance was provided by I. Harrison and P. Donahoe,
3 1. C,, 13@(H), 14o(H) tricyclic terpane (22R) and the artwork produced by Ms. C. Jeans (NRG).
32. C,, 13B(H),14a(H) tricyclic terpane (22s) D. Ariztegui, S. Bernasconi and J. McKenzie (ETH,
33. C,, 13B(H),14o(H) tricyclic terpane (22R) Zurich) are thanked for their help in obtaining the
34. C,, 13B(H),14u(H) tricyclic terpane (22s)
35. C,, 13B(H), 141x(H) tricyclic terpane (22R)
stable isotopic data. We are also grateful to R. Tyson
(NRG) for invaluable discussions regarding this
Tetracyclic terpane: work, and to the reviewers of the manuscript for
36. C,, tetracyclic terpane
their constructive comments. (RA)

Sterunes:

37. C,, k(H), 14E(H), 17P(H) sterane (20R) References


38. Ci, 5a(H), 14B(H),17P(H) sterane (20s)
39. C,, ~IX(H),~~B(H),~~B(H) sterane (20R) Aquino Neto, F.R., Triguis, D.A., Azevedo, D.A., Rodrigues, R.
40. C,, 5a(H),14P(H),l7P(H) sterane (20s) and Simoneit, B.R.T., 1992. Organic geochemistry of geo-
41. C,, 5c~(H),14B(H),l7B(H) sterane (20R) graphically unrelated Tamunites. Org. Geochem., 18: 791-
42. C,, 5a(H),l4B(H),17B(H) sterane (20s) 803.
Arthur, M.A., Dean, W.E. and Claypool, GE., 1985. Anomalous
J. Perez-Infante et d/Chemical Geology 130 (1996) 271-288 287

13C enrichment in modem marine organic carbon. Nature Haq, B.U., Hardenbol, J. and Vail, P.R., 1987. Chronology of
(London), 315: 216-218. fluctuating sea levels since the Triassic. Science, 235: 1156-
Arthur, M.A., Schlanger, SO. and Jenkyns H.C., 1987. The 1166.
Cenomanian-Turonian Oceanic Anoxic Event, II. Paleoceano- Hay, W.W., 1985. Paleoceanography of the Venezuelan Basin. 1st
graphic controls on organic matter production and preserva- Geol. Conf., Geol. Sot. Trinidad-Tobago, Proc., pp. 302-307.
tion. In: J. Brooks and A.J. Fleet (Editors), Marine Petroleum Hayes, J.M., Popp, B.N., Takigiku, R. and Johnson, M.W., 1989.
Source Rocks. Geol. Sot. London, Spec. Publ., 26: 401-420. An isotopic study of biogecchemical relationships between
Arthur, M.A., Jenkyns, H.C., Brumsack, H.J. and Schlanger, S.O., carbonates and organic carbon in the Greenhorn Formation.
1990. Stratigraphy, geochemistry and paleoceanography of Geochim. Cosmochim. Acta, 53: 2961-2972.
organic carbon-rich Cretaceous sequences. In: R.N. Ginsburg Hilbrccht, H. and Hoefs, J., 1986. Geochemical and paleontologi-
and B. Beaudin (Editors), Cretaceous Resources, Events and cal studies of the 613C anomaly in boreal and north Tethyan
Rhythms. Kluwer, Dordrecht, pp. 75-l 19. Cenomanian-Turonian sediments in Germany and adjacent
Berger, W.H., 1970. Planktonic foraminifera: selective solution areas. Palaeogeogr., Palaeoclimatol., Palaeoecol., 53: 169- 189.
and the lysocline. Mar. Geol., 8: 111-138. Hilbrecht, H., Hubberten, H. and Oberhansli, H., 1992. Biogeog-
Bralower, T.J., 1988. Calcareous nannofossil biostratigraphy and raphy of planktonic foraminifera and regional carbon isotope
assemblages of the Cenomanian-Turonian boundary interval. variations: productivity and water masses in Late Cretaceous
Paleoceanography, 3: 275-3 16. Europe. Palaeogeogr., Palaeoclimatol., Palaeoecol., 92: 407-
Bremner, J.M., 1983. Biogenic sediments in the South West 421.
African (Namibian) continental margin. In: J. Thiede and E. Hollander, D., 1989. Carbon and isotopic cycling and organic
Suess (Editors), Coastal Upwelling: Its Sediment Record, Part geochemistry of eutrophic Lake Greifen: Implications for
B: Sedimentary Records of Ancient Coastal Upwellin. NATO preservation and accumulation of ancient organic carbon rich
(N. Atlantic Treaty Org.) Conf. Ser. IV, lob, Plenum, New sediments. Ph.D. Thesis, ETH (Eidgeniissische Technische
York, N.Y., pp. 73-103. Hochschule), Zurich.
Craig, H., 1957. isotopic standards r carbon and oxygen and Hut, A.Y., 1988. Aspects of depositional processes of organic
correction factors for mass-spectrometric analysis of carbon matter in sedimentary basins. Org. Geochem., 13: 263-272.
dioxide. Geochim. Cosmochim. Acta, 12: 133-149. James, K.H. 1990. The Venezuelan hydrocarbon habitat. In: J.
Davis, J.C., 1986. Statistics and Data Analysis in Geology. Wiley, Brooks (Editor), Classic Petroleum Provinces. Geol. Sot. Lon-
New York, N.Y., 2nd ed., 646 pp. don, Spec. Publ., 50: 9-35.
Einsele, G. and Wiedmann, J., 1982. Turonian black shales in the Jenkyns, H.C., 1980. Cretaceous anoxic events: from continents to
Moroccan coastal basins: first upwelling in the Atlantic Ocean. oceans. J. Geol. Sot. London, 137: 171-88.
In: U. von Rad, K. Hinz, M. Sarntheim and E. Seibold Jenkyns, H.C., Gale, A.S. and Corfield, R.M., 1994. Carbon- and
(Editors), Geology of the Northwest African Continental Mar- oxygen-isotope stratigraphy of the English Chalk and Italian
gin. Springer, Berlin, pp. 396-414. Scaglia and its palaeoclimatic significance. Geol. Mag., 131:
Erikson, J.P. and Pindell, J.L., 1993. Analysis of subsidence in l-34.
northeastern Venezuela as a discriminator of tectonic models Kauffman, E.G., 1984. The fabric of Cretaceous marine extinc-
for northern South America. Geology, 21: 945-948. tions. In: W.A. Berggren and J.A. Van Couvering (Editors),
Farrimond, P., Eglinton, G., Brassell, S.C. and Jenkyns, H.C., Catastrophes and Earth History: the New Uniformitarianism.
1990. The Cenomanian/Turonian anoxic event in Europe: an Princeton University Press, Princeton, N.J., pp. 151-246.
organic geochemical study. Mar. Pet. Geol., 7: 75-89. Koutsoukos, E.A.M. and Hart, M.B., 1990. Radiolarians and
Farrimond, P., Stoddart, D.P. and Jenkyns, H.C., 1994. An or- diatoms from the mid-Cretaceous successions of the Sergipe
ganic geochemical profile of the Toarcian anoxic event in Basin, northeastern Brazil: palaeoceanographic assessment. J.
notthem Italy. Chem. Geol., 111: 17-33. Micropalaeontol., 9: 45-64.
Ford, A. and Houbolt, J.J., 1963. The Microfacies of the Creta- Kuhnt, W., Herbin, J.P., Thurow, J. and Wicdmann, J., 1990.
ceous of Western Venezuela. Brill, Leiden, 109 pp. Distribution of Cenomanian-Turonian organic facies in the
Gale, A.S., Jenkyns, H.C., Kennedy, W.J. and Cortield, R.M., Western Mediterranean and along the adjacent Atlantic mar-
1993. Chemostratigraphy versus biostratigraphy: data from gin. In: A.Y. Hut (Editor), Deposition of Organic facies. Am.
around the Cenomaniar-Turonian boundary. J. Geol. Sot., Assoc. Pet. Geol. Stud. Geol., 30: 133-160.
London, 150: 29-32. Macellati, C.E. 1988. Cretaceous paleogeography and depositional
Gawthorpe, R., Hunt, D., Taylor, A. and Underhill, J., 1994. cycles of western South America. J. S. Am. Earth Sci., 1:
Course Notes of NERC Sequence Stratigraphy Workshop. 373-418.
Dep. Geol., Univ. of Manchester, Manchester. Macellari, C.E. and De Vries, T.J., 1987. Late Cretaceous up-
Gonzalez de Juana, C., Iturrilde de Arocena, J. and Picard, X., welling and anoxic sedimentation in northwestern South
1980. Geologia de Venezuela y de sus Cuencas Petroliferas. America. Palaeogeogr., Palaeoclimatol., Palaeoecol., 59: 279-
Foninves, Caracas, 103 1 pp. 292.
Hancock, J.M., 1993. Sea-level changes around the Martinez, J.I. and Hemandez, R., 1992. Evolution and drowning
Cenomanian-Turonian boundary. Cretaceous Res., 14: 553- of the Late Cretaceous Venezuelan carbonate platform. J. S.
562. Am. Earth Sci., 5: 197-210.
288 .I. Perez-infante et ul./Chemicul Geology I30 (1996) 271-288

McCaffrey, M.A., Simoneit, B.R.T., Aquino Neto, F.R. and Part A: Responses of the Sedimentary regime to present
Moldowan, J.M., 1994. Functionalized biological precursors Coastal Upwelling. NATO (N. Atlantic Treaty Org.) Conf.
of tricyclic terpanes: information from sulfur-bound biomark- Ser., IV, IOa, Plenum, New York, N.Y., pp. l-10.
ers in a Permian Tasmanite. Org. Geochem., 21: 481-488. Talukdar, S., Gallango, 0. and Chin-A-Lien, M., 1985. Genera-
Mello, M.R., Koutsoukos, E.A.M., Hart, M.B., Brassell, SC. and tion and migration of hydrocarbons in the Maracaibo Basin,
Maxwell, J.R., 1989. Late Cretaceous anoxic events in the Venezuela: an integrated basin study. Org. Geochem., 10:
Brazilian continental margin. Org. Geochem., 14: 529-542. 261-279.
Molina Cruz, A., 1984. Radiolaria as indicators of upwelling Tissot, B.P. and Welte, D.H.. 1984. Petroleum Formation and
processes: the Peruvian connection. Mar. Micropaleontol., 9: Occurrence. Springer, Berlin, 699 pp.
53-75. Tribovillard, N.P., Stephan, J.F., Manivit, H., Reyre, Y., Cotillon,
Morse, J.W. and Mackenzie, F.T., 1990. Geochemistry of Sedi- P. and Jautee, E., 1991. Cretaceous black shales of Venezue-
mentary Carbonates. Elsevier, Amsterdam, 707 pp. lan Andes: preliminary results on stratigraphy and paleoenvi-
Nederbragt, A., 1991. Late Cretaceous biostratigraphy and devel- ronmental interpretations. Palaeogeogr., PaIaeocIimatol.,
opment of Heterohelicidae (planktic foraminifera). Micropale- Palaeoecol., 81: 313-321.
ontology, 37: 329-372. Tyson, R.V., 1995. Sedimentary Organic Matter. Chapman and
Noble, R., Alexander, R. and Kagi, RI., 1984. The Occurrence of Hall, London, 615 pp.
bisnorhopane, trisnorhopane and 25.norhopanes as free hydro- Vahrenkamp, V.C., Franssen, R.C., Grotsch, J. and Muiioz, P.J.,
carbons in some Australian shales. Org. Geochem., 8: 171~ 1993. Maracaibo Platform (Aptian-Albian), Northwestern
176. Venezuela. In: J.A. Toni, R. Scott and J.P. Masse (Editors),
Peters, K.E. and Moldowan, J.M., 1993. The Biomarker Guide: Cretaceous Carbonate Platforms. Am. Assoc. Pet. Geol. Mem.,
Interpreting Molecular Fossils in Petroleum and Ancient Sedi- 56: 25-33.
ments. Prentice Hall, London, 346 pp. Volkman, J.K., 1988. Biological marker compounds as indicators
Popp, B.N., Takigiku, R., Hayes, J.M., Louda, J.W. and Baker, of the depositional environments of petroleum source rocks.
E.W., 1989. The post-Paleozoic chronology and mechanism of In: A.J. Fleet, K. Kelts and M.R. Talbot (Editors), Lacustrine
“C depletion in primary marine organic matter. Am. J. Sci., Petroleum Source Rocks. Geol. Sot. London, Spec. Publ., 40:
289: 436-454. 103-122.
Ricken, W., 1994. Complex rhythmic sedimentation related to Volkman, J.K., Banks, M.R., Denwer, K. and Aquino Neto, F.R.,
third-order sea level variations. In: P.L. Boer and D.G. Smith 1989. Biomarker composition and depositional setting of tas-
(Editors), Orbital Forcing and Cyclic Sequences. Spec. Publ. manite oil shale from Northern Tasmania, Australia. EAOG
hit. Assoc. Sedimentol., 19: 167-193. (Eur. Assoc. Org. Geochem.) 14th Int. Meet., Paris, Abstr.
Ricken, W. and Eder, W., 1991. Diagenetic modification of 168.
calcareous beds - an overview. In: G. Einsele, W. Ricken Welling, L.A., Pisias, N.G. and Roelofs, A.K., 1992. Radiolarian
and A. Seilacher (Editors), Cycles and Events in Stratigraphy. microfauna in the northern California Current System: indica-
Springer, Berlin, pp. 430-449. tors of multiple processes controlling productivity. In: C.P.
Schlanger, S.O. and Jenkyns, H.C., 1976. Cretaceous anoxic Summerhayes, W.L. Prell and K.C. Emeis (Editors), Up-
events: causes and consequences. Geol. Mijnbouw, 55: 1799 welling Systems: Evolution Since the Early Miocene. Geol.
184. Sot. London, Spec. PubI., 64: 177- 195.
Schlanger, S.O., Arthur, M.A., Jenkyns, H.C. and Scholle, P.A., Wignall, P.B. and Maynard, J.R., 1993. The sequence stratigraphy
1987. The Cenomaniat-Turonian oceanic anoxic event, 1. of transgressive black shales. In: B.J. Katz and L. Pratt
Stratigraphy and distribution of organic-rich beds and the (Editors), Source Rocks in a Sequence Stratigraphy Frame-
marine 813C excursion. In: J. Brooks and J. Fleet (Editors), work. Am. Assoc. Pet. Geol., Studies in Geology, 37: 35-47.
Marine Petroleum Source Rocks. Geol. Sot. London, Spec. Wilson, J.L., 1975. Carbonate Facies in Geologic History.
PubI., 26: 371-399. Springer, New York, N.Y., 471 pp.
Scholle, P.A. and Arthur, M.A., 1980. Carbon isotope fluctuations Zambrano, E., Vazquez, E., Duval, B., Latreille, M. and
in Cretaceous pelagic limestones: potential stratigraphic and Coffmieres, B., 1971. Sintesis paleogeografica y petrolera del
petroleum exploration tool. Am. Assoc. Pet. Geol. Bull., 64: occidente de Venezuela. Mem. IV Congr. GeoI. Venezolano,
67-87. Caracas, pp. 483-552.
Simoneit, B.R.T., Leif, R.N., Aquino Neto, F.R., Azevedo, D.A., Zumberge, J.E. 1984. Source rocks of the La Luna Formation
Pinto, A.C. and Albrecht, P., 1990. On the presence of tr- (Upper Cretaceous) in the Middle Magdalena valley, Colom-
cyclic terpane hydrocarbons in Permian tasmanite algae. bia. In: J.G. Palacas (Editor), Petroleum Geochemistry and
Naturwissenschaften, 77: 380-383. Source Rock Potential of Carbonate Rocks. Am. Assoc. Pet.
Suess, E. and Thiede, J., 1983. Introduction. In: E. Suess and J. GeoI. Stud. Geol., 18: 127-133.
Thiede (Editors), Coastal Upwelling: Its Sedimentary Record,

You might also like