You are on page 1of 199

pal273-3 cover 2/16/09 4:52 PM Page 2

www.elsevier.com/locate/pgeola

bra
Editors-in-Chief le t

in
150
D.J. Bottjer P. Kershaw

C
Department of Earth Sciences School of Geography and
University of Southern California Environmental Science

g
Los Angeles, California 90089-0740 Monash University


USA P.O. Box 11A
Tel.: 1 213 740 6100
Fax: 1 213 740 8801
Clayton VIC. 3800
Australia
Proceedings of the


E-mail: dbottjer@usc.edu Tel.: 61 3 9905 2910
ea

Y
Fax: 61 3 9905 2948
T. Corrège E-mail: peter.kershaw@arts.monash.edu.au rs
Universite Bordeaux 1
UMR CNRS 5805 EPOC
Avenue de Facultés, 33405 Talence Cedex
France
Tel.: 33 5 4000 8731
F. Surlyk
Institute of Geography and Geology
University of Copenhagen
Øster Voldgade 10
Geologists’ Association
Fax: 33 5 5684 0848 DK 1350 Copenhagen K
E-mail: t.correge@epoc.u-bordeaux1.fr Denmark
Tel.: 45 35 32 24 53
Fax: 45 33 14 83 22
E-mail: Finns@geol.ku.dk
STARTING 2009 PUBLISHED BY ELSEVIER
Editorial Board

J. Barron, Menlo Park, CA L.J. Flynn, Cambridge, MA J.I. Martinez, Medelin


E.J. Barron, Austin, TX S. Fritz, Lincoln, NE R.B. Pearce, Southampton
D. Beerling, Sheffield F.T. Fürsich, Erlangen I. Premoli Silva, Milano
H. Behling, Göttingen
J. Giraudeau, Talence G. Retallack, Eugene, OR The Proceedings of the Geologists' Association is an international

BENEFITS
M.J. Benton, Bristol A.C. Scott, Egham, Surrey
M.I. Bird, Fife M.J. Head, St. Catharines, ON geosciences journal that was founded in 1859 and publishes research and
S. Hesselbo, Oxford P. Swart, Miami, FL
H.J.B. Birks, Bergen M.R. Talbot, Bergen review papers on all aspects of Earth Science. In particular, papers will focus
I. Boomer, Birmingham A. Holbourn, Kiel TO AUTHORS INCLUDE:
J.C. Braga, Granada T.C. Johnson, Duluth, MN
E. Thomas, New Haven, CT on the geology of the region around the southwestern margin of the North
M.R. Walter, North Ryde, N.S.W. • Easy online submission at:
C.E. Brett, Cincinnati, OH K. Kaiho, Sendai P. Wang, Shanghai
Sea basin, giving especial attention to the Mesozoic, Cenozoic and
R.G. Bromley, Copenhagen S. Kershaw, Uxbridge, Middlesex http://ees.elsevier.com/pgeola
J.H. Cann, Adelaide P.B. Wignall, Leeds Quaternary rocks and landforms that characterise this region.
D.B. Lazarus, Berlin C. Whitlock, Bozeman, MT
T.J. Crowley, Edinburgh • Submissions are free of charge
P. De Deckker, Canberra, A.C.T. J. Lee-Thorp, Rondebosch Y. Yokoyama, Tokyo Papers on applied, historical and educational aspects of Earth Science are
Z. Ding, Beijing A. Longinelli, Parma G. Zanchetta, Pisa • No page charges
welcome and the PGA will publish reports on field meetings, held by the
N. Eyles, Scarborough, ON M.O. Manceñido, La Plata M. Zhu, Nanjing
• Fast maximum online visibility Association, that cover new findings and interpretations. The journal will
Scope of the journal
Palaeogeography, Palaeoclimatology, Palaeoecology is an international medium for the publication of high quality and multidisciplinary original studies and comprehensive (within three weeks after
reviews in the field of palaeo-environmental geology. The journal aims at bringing together data with global implications from the research in the many different disciplines
publish, as appropriate, Special Issues on topics of scientific importance,
involved in palaeo-environmental investigations.
acceptance)
and will welcome substantive reviews where
Publication information • 30% discount on most
Palaeogeography, Palaeoclimatology, Palaeoecology (ISSN 0031-0182). For 2009, volumes 271– 284 are scheduled for publication. A combined subscription to Palaeogeography,
these cover topics of importance
Elsevier books
Palaeoclimatology, Palaeoecology and Global and Planetary Change (Vols. 65– 69) at reduced rate is available. Subscription prices are available upon request from the Publisher
or from the Regional Sales Office nearest you or from this journal’s website (http://www.elsevier.com/locate/palaeo). Further information is available on this journal and other
to Earth Science.
Elsevier products through Elsevier’s website (http://www.elsevier.com). Subscriptions are accepted on a prepaid basis only and are entered on a calendar year basis. Issues are sent
by standard mail (surface within Europe, air delivery outside Europe). Priority rates are available upon request. Claims for missing issues should be made within six months of
the date of dispatch.

Orders, claims, and journal enquiries: please contact the Regional Sales Office nearest you:
St. Louis: Elsevier, Customer Service Department, 11830 Westline Industrial Drive, St. Louis, MO 63146, USA; phone: (877) 8397126 [toll free within the USA]; (+1) (314) 4537076
[outside the USA]; fax: (+1) (314) 5235153; e-mail: JournalCustomerService-usa@elsevier.com
Amsterdam: Elsevier, Customer Service Department, PO Box 211, 1000 AE Amsterdam, The Netherlands; phone: (+31) (20) 4853757; fax: (+31) (20) 4853432; e-mail:
JournalsCustomerServiceEMEA@elsevier.com
Tokyo: Elsevier, Customer Service Department, 4F Higashi-Azabu, 1-Chome Bldg, 1-9-15 Higashi-Azabu, Minato-ku, Tokyo 106-0044, Japan; phone: (+81) (3) 5561 5037; fax:
(+81) (3) 5561 5047; e-mail: JournalsCustomerServiceJapan@elsevier.com
Singapore: Elsevier, Customer Service Department, 3 Killiney Road, #08-01 Winsland House I, Singapore 239519; phone: (+65) 63490222; fax: (+65) 67331510; e-mail:
JournalsCustomerServiceAPAC@elsevier.com
Editor-in-Chief:
Advertising information
J. Rose
Advertising orders and enquiries can be sent to: James Kenney, Elsevier Ltd., 32 Jamestown Road, London NW1 7BY, UK; phone: (+44) 207 424 4216; fax: (+44) 1865 853 136; University of London, Royal Holloway,
e-mail: j.kenney@elsevier.com. Customers in the US and Canada can also contact: Mr Tino DeCarlo, Advertising Department, Elsevier Inc., 360 Park Avenue South, New York, NY Egham Hill, UK
10010-1710, USA; phone: (+1) (212) 633 3815; fax: (+1) (212) 633 3820; e-mail: t.decarlo@elsevier.com

USA mailing notice: Palaeogeography, Palaeoclimatology, Palaeoecology (ISSN 0031-0182) is published semimonthly by Elsevier B.V. (Elsevier B.V., Radarweg 29, 1043 NX,
Amsterdam, The Netherlands). Periodical postage paid at Rahway NJ and additional mailing offices.
Editors:
USA POSTMASTERS: Send change of address to Palaeogeography, Palaeoclimatology, Palaeoecology, Elsevier, Customer Service Department, 11830 Westline Industrial Drive, St. Louis,
MO 63146, USA.
D.J. Horne
AIRFREIGHT AND MAILING in the USA by Mercury International Limited, 365, Blair Road, Avenel, NJ 07001. Queen Mary University of London, UK
The paper used in this publication meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper) J.H. Powell
Printed by Polestar Wheatons Ltd, Exeter, UK
British Geological Survey,
Natural Environment Research Council,
Nottingham, Keyworth, UK
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) iii–iv

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Special Issue

Organic-carbon-rich sediments through the Phanerozoic: Processes, progress, and perspectives

Edited by
A. Negri a, A. Ferretti b, T. Wagner c, P.A. Meyers d

a
Dipartimento di Scienze del Mare, Università Politecnica delle Marche, Via Brecce Bianche, 60131 Ancona, Italy
b
Dipartimento di Scienze della Terra, Università degli Studi di Modena e Reggio Emilia, Largo S. Eufemia, 41100 Modena, Italy
c
Department of Civil Engineering and Geosciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
d
Department of Geological Sciences, The University of Michigan, Ann Arbor, Michigan, 48109-1005, USA

CONTENTS

Publisher’s note . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v

Prefaces
Organic-carbon-rich sediments through the Phanerozoic: Processes, progress, and perspectives
A. Negri, A. Ferretti, T. Wagner and P.A. Meyers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 213
Phanerozoic organic-carbon-rich marine sediments: Overview and future research challenges
A. Negri, A. Ferretti, T. Wagner and P.A. Meyers . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 218

Research Papers
Transport and depositional process of soil organic matter during wet and dry storms on the Têt inner shelf (NW Mediterranean)
J.-H. Kim, R. Buscail, F. Bourrin, A. Palanques, J.S. Sinninghe Damsté, J. Bonnin and S. Schouten . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 228
Geochemical evidence for high-resolution variations during deposition of the Holocene S1 sapropel on the Cretan Ridge, Eastern Mediterranean
G. Gennari, F. Tamburini, D. Ariztegui, I. Hajdas and S. Spezzaferri . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 239
Role of sea-level forced sedimentary processes on the distribution of organic carbon-rich marine sediments: A review of the Late
Quaternary sapropels in the Mediterranean Sea
R. Capozzi and A. Negri. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 249
Benthic environmental changes in the Eastern Mediterranean Sea during sapropel S5 deposition
C. Morigi . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 258
Organic geochemistry and paleoenvironment of the Early Eocene ‘‘Pesciara di Bolca’’ Konservat-Lagerstätte, Italy
L. Schwark, A. Ferretti, C.A. Papazzoni and E. Trevisani . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 272
Geochemical environment of the Coniacian–Santonian western tropical Atlantic at Demerara Rise
C. März, B. Beckmann, C. Franke, C. Vogt, T. Wagner and S. Kasten . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 286
Paleo-redox conditions during OAE 2 reflected in Demerara Rise sediment geochemistry (ODP Leg 207)
A. Hetzel, M.E. Böttcher, U.G. Wortmann and H.-J. Brumsack . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 302
Abstract of ‘‘Climate-Ocean coupling off North-West Africa during the Lower Albian: The Oceanic Anoxic Event 1b’’
P. Hofmann, I. Stüsser, T. Wagner, S. Schouten and J.S. Sinninghe Damsté . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 329
Temperature controlled deposition of early Cretaceous (Barremian–early Aptian) black shales in an epicontinental sea
J. Mutterlose, S. Pauly and T. Steuber . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 330
Late Pliensbachian–Early Toarcian (Early Jurassic) environmental changes in an epicontinental basin of NW Europe (Causses area, central
France): A micropaleontological and geochemical approach
S. Mailliot, E. Mattioli, A. Bartolini, F. Baudin, B. Pittet and J. Guex . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 346
Abstract of ‘‘The role of sea-level change and marine anoxia in the Frasnian–Famennian (Late Devonian) mass extinction’’
D.P.G. Bond and P.B. Wignall . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 365
Black shale deposition in an Upper Ordovician–Silurian permanently stratified, peri-glacial basin, southern Jordan
H.A. Armstrong, G.D. Abbott, B.R. Turner, I.M. Makhlouf, A.B. Muhammad, N. Pedentchouk and H. Peters . . . . . . . . . . . . . . . . . . . . . . . 368

doi:10.1016/S0031-0182(09)00059-5
iv Special Issue

Palynology, organic geochemistry and carbon isotope analysis of a latest Ordovician through Silurian clastic succession from borehole
Tt1, Ghadamis Basin, southern Tunisia, North Africa: Palaeoenvironmental interpretation
M. Vecoli, A. Riboulleau and G.J.M. Versteegh . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 378
Organic-carbon deposition and coastal upwelling at mid-latitude during the Upper Ordovician (Late Katian): A case study from the
Welsh Basin, UK
T.J. Challands, H.A. Armstrong, D.P. Maloney, J.R. Davies, D. Wilson and A.W. Owen . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 395
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) v

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Publisher's note

We regret that owing to an oversight on our part two papers P. Hofmann, I. Stüsser, T. Wagner, S. Schouten, J.S. Sinninghe Damsté,
handled by the Guest Editors and accepted for inclusion in the Climate-Ocean coupling off North-West Africa during the Lower Albian:
special issue on Organic-carbon-rich sediments through the Phaner- The Oceanic Anoxic Event 1b, Palaeogeography, Palaeoclimatology,
ozoic: Processes, progress, and perspectives have been published Palaeoecology 262, 2008, 157-165, doi:10.1016/j.palaeo.2008.02.014.
already:
We apologise to the authors, Guest Editors and readers for the
D.P.G. Bond and P.B. Wignall, The role of sea-level change and inconvenience caused by this. In the special issue abstracts of these
marine anoxia in the Frasnian-Famennian (Late Devonian) mass papers are included, to indicate their place in the sequence of papers.
extinction, Palaeogeography, Palaeoclimatology, Palaeoecology 263,
2008, 107-118, doi:10.1016/j.palaeo.2008.02.015. The Publisher

doi:10.1016/S0031-0182(09)00054-6
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 213–217

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Preface

Organic-carbon-rich sediments through the Phanerozoic: Processes, progress,


and perspectives
A. Negri a,⁎, A. Ferretti b, T. Wagner c, P.A. Meyers d
a
Dipartimento di Scienze del Mare, Università Politecnica delle Marche, Via Brecce Bianche, 60131 Ancona, Italy
b
Dipartimento di Scienze della Terra, Università degli Studi di Modena e Reggio Emilia, Largo S. Eufemia, 41100 Modena, Italy
c
School of Civil Engineering and Geosciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
d
Department of Geological Sciences, The University of Michigan, Ann Arbor, Michigan 48109-1005 USA

a r t i c l e i n f o

Article history:
Received 3 October 2008
Received in revised form 4 November 2008
Accepted 13 November 2008

1. Introduction by “Causes and Consequences of Marine Organic Carbon Burial Through


Time” (Negri et al., 2006). These two volumes brought together the
This Special Issue is comprised of two related components. First, sapropel and black shale communities that had previously studied these
we present a summary of the occurrences of organic-carbon-rich organic-carbon-rich sequences separately. Moreover, their goal was to
sedimentary sequences in the Phanerozoic geological record and our highlight the similarities and differences of these two kinds of interesting
overview of the evolution and current status of understanding of how sediments, pointing to mergers of the knowledge acquired separately (e.g.,
they accumulated. We note especially some of the new advances in the Meyers, 2006).
kinds of palaeoceanographic proxies that can be used and in the more With the current Special Issue we expand the coverage of geological
refined explanations that have emerged from these advances. The time to include the Palaeozoic Era to explore the processes involved in
second and greater part of the Special Issue consists of fourteen deposition of organic-carbon-rich sequences in this huge interval of
research papers that provide examples of the more detailed analyses deep time (291 Ma). Investigators of Palaeozoic settings must cope with
of organic-carbon-rich sedimentary sequences that can be achieved the problems of lower time resolution, seldom continuous outcrops, and
with modern techniques and that can be elegantly interpreted by general lack of undisturbed deep sea sediment sequences. Study of all
specialists in these techniques. palaeoceanographic processes is necessarily mediated by the available
The stimulus for this Special Issue emerged from a very successful time resolution, and in the Palaeozoic the length of individual biozones
topical session presented at the European Geophysical Union Meeting in is generally on the order of millions of years, which is in the same range
Vienna, Austria, in April 2007. The very well attended (80–100 people) as third-order sea-level changes. Thus, an important question in
session addressed recent research advances in “Organic-carbon rich Palaeozoic sequences is whether episodes occur at different scales or
sediment through time: Past present and future, ocean and climate belong to cycles of diverse order. Nevertheless, the study of these very
feedback” that included a large number of excellent oral and poster old sediments cannot be neglected because they represent intervals of
presentations. This session was the continuation of a project started in time that are characterized by cycling of tremendous amounts of CO2,
2000 that has already yielded two well-received special issues of Palaeo- which is one of the main factors responsible for profound changes in the
geography, Palaeoclimatology, Palaeoecology. Both special issues present Earth–Ocean–Atmosphere system and whose increase is central to
collections of contributions that deal with organic-carbon-rich marine modern climate changes.
sequences in different temporal windows, and both simultaneously Within this broad framework, a combined geochemical, palaeontolo-
emphasize new issues in research and reviews of existing concepts and gical, and biological approach can address prominent examples in Earth
models. The current volume forms a trilogy launched by the Special Issue history to assess the mechanisms operating during natural rapid global
“Paleoclimatic and Paleoceanographic Records in Mediterranean Sapro- change either in greenhouse (e.g., the Cretaceous) or icehouse (e.g.,
pels and Mesozoic Black Shales” (Meyers and Negri, 2003) and continued Pliocene–Pleistocene) conditions. Better integration of observations of
past and current climatic conditions can considerably enhance our global
understanding of basic aspects of the ocean carbon system and associated
⁎ Corresponding author. climate change. In addition, a growing recognition exists that the concepts
E-mail address: a.negri@univpm.it (A. Negri). and models that have been proposed to explain how organic-carbon-rich

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.11.016
214 Preface

Fig. 1. Organic carbon-rich sediments through time (Photo credits: Petr Štorch, A.N., T.W. and ISMAR CNR Bologna). A: Holocene sapropel S1 in the Ionian Basin. B: The most
spectacular outcrop of pelagic Cretaceous sediments in the Umbria Marche Basin of Italy: the Vispi Quarry in the Contessa Valley, near Gubbio. Sediments span 80 My and belong to
four formations: Maiolica, Marne a Fucoidi, Scaglia Bianca, and Scaglia Rossa. The Marne a Fucoidi Fm. (MaF) shows evident cyclic alternations of light carbonate rich marls and darker
shales related to Milankovitch orbital perturbations. Arrow indicates the Selli Level (OAE1a, late early Aptian). C: Coastal exposure of upper Cenomanian marl-black shale sequences
at Mohammed Plage, North of Tarfaya, southern Morocco. Alternations of light colour carbonate-rich marls and dark organic carbon-rich black shale document regular changes in
environmental and climatic conditions that occurred at orbital frequencies. D: Close-up of organic carbon-rich black shales showing mm-scale lamination indicative of extreme
oxygen depletion at the time of deposition. E: Prague Basin, Kosov Quarry, Silurian black shales of the Motol Formation (Wenlock, ramosus-ellesae and lundgreni graptolite biozones).

marine sediments are deposited merit reconsideration. Are there other of information about possible future oceanic responses to a progressively
mechanisms for organic carbon preservation that are viable alternatives to warming world. Recent observations in the equatorial Atlantic and Indian
the classical anoxia vs. productivity models? oceans that support decreasing oxygenation of the tropical ocean
The challenge therefore is to understand if the scientific community, (Stramma et al., 2008) highlight the urgency of this topic and emphasize
notwithstanding their respective specialization in different time windows, immediate demand for integrated research strategies.
can come to a common language and a common system that can
disentangle critical elements of the earth history puzzle. This new 1.1. Note
knowledge will directly impact our ability to make realistic future
projections beyond the human-scale time horizon and to assess potential Two papers that should have been included in this Special Issue
carbon management scenarios, and it will represent a fundamental source were by mistake published before (Bond and Wignall, 2008; Hofmann
Preface 215

Table 1
Main features of the papers included in this special issue

et al., 2008). In their place in this issue (extended) abstracts are in- sapropel S1. They identify fluctuations in major and minor chemical
cluded (Bond and Wignall, 2009-this issue; Hofmann et al., 2009-this element distributions that reveal high-frequency cyclicities that
issue). correspond to millennial and centennial-scale solar cycles. This
correspondence suggests that sapropel deposition is ultimately linked
2. Organization and content of the research papers to global processes, despite being confined to the Mediterranean
basin.
The fourteen contributions to this Special Issue on organic-carbon- Capozzi and Negri (2009-this issue) investigate the role of sea-
rich sedimentation are organized by increasing geological age (see level forced sedimentary processes and their impact on distribution of
Table 1). Although this order is the reverse of how the sequences were the Late Quaternary sapropels in the Mediterranean Sea. They
deposited, we elected it for two reasons. First, recent and near-recent examine published data on the occurrence of the complete succession
sequences are more easily studied in detail than older and often less- of the S1-S5 sapropels and then focus on two stratigraphic sequences
complete sequences. Second, the Pliocene–Pleistocene sapropels of in the central Adriatic shelf and Middle Adriatic Depression. They find
the Mediterranean Sea are increasingly being considered as near- that sapropels S5 and S1 occur during periods of sea level rise
modern analogs of Mesozoic and Palaeozoic black shales. Under- recorded by the Transgressive System Tracts (Marine Isotopic Stage,
standing the palaeoceanographic and palaeoclimatic processes that MIS, 5 and 1 respectively). In contrast, sapropels S4 and S3 were
led to sapropel deposition therefore can improve understanding of deposited during the Highstand System Tracts that developed
how older organic-carbon-rich sequences were deposited. The range throughout MIS 5 during the late Pleistocene sequence 1. Sapropel
of geological ages that the fourteen contributions span is from the S2 deposition, however, occurred during the warm MIS 3.3 and
modern to the Ordovician. The papers are consequently grouped into follows the MIS 4 sea level drop when an enhanced sediment supply is
Cenozoic, Mesozoic, and Palaeozoic units. recorded during the late highstand and the falling sea level stage of
The Cenozoic unit of this Special Issue begins with a study by Kim sequence 1. They also consider the possible influence of sea level
et al. (2009-this issue) who investigate the modern processes involved lowering on Mediterranean deep water formation and conclude that
with the transport and deposition of soil organic matter during wet although favourable conditions could be triggered by sea level
and dry storms on the Têt inner shelf of the northwest Mediterranean oscillation, greater productivity is likely the main cause of the
Sea to show that these events act as a bypass corridor for delivery of sedimentation of sapropels.
land-derived organic matter to the seafloor. Morigi (2009-this issue) reconstructs the environmental changes
Next, Gennari et al. (2009-this issue) present the results of their in the Eastern Mediterranean Sea during sapropel S5 deposition by
high resolution study of the geochemical signatures of the Holocene means of a high resolution benthic foraminiferal analysis in multiple
216 Preface

sediment cores. The study documents that at relatively shallow depths drawdown of the seawater trace metal reservoir, based on the general
in the bathyal environment the sea-floor was partially ventilated and observation that hypoxic to euxinic environments form a sink for trace
that re-oxygenation had considerably increased during the late phase metals (e.g., Brumsack, 2006).
of S5 deposition. However, at deeper locations benthic abundance and Hofmann et al. (2008, 2009-this issue) address aspects of climate–
diversity strongly decrease during sapropel S5, and microfauna even ocean coupling off northwest Africa during the Lower Albian OAE1b.
disappear in some levels, suggesting the establishment and main- They present high resolution records of organic and inorganic
tenance of stagnant and anoxic conditions at the sea floor until the geochemical proxies from DSDP Site 545 on the Mazagan Plateau
end of S5 deposition. In the deepest part of the basin, gradual that document a complex and rapid environmental change for
repopulation of the benthic foraminiferal community at the top of the onshore continental climate and the offshore upwelling system
S5 layer indicates a relatively slow bottom re-oxygenation. In the associated with OAE1b. TEX86-based SST estimates reveal an abrupt
southernmost site, the benthic foraminifera assemblage records a rise of ~3 °C concurrent with a more than two-fold increase in ac-
short oxygenation pulse during the S5 deposition that is linked to a cumulation rates for organic-carbon and siliciclastic sediment
short cold spell. Results show that the evolution of the dysoxic–anoxic components, with an influx of overall finer grained sediment as
conditions as well as the re-oxygenation pattern at the end of the inferred from Si/Al and Zr/Al ratios. This new set of data supports the
stagnant period was strongly dependent on the basin morphology. conclusion that warming associated with OAE1b resulted in an
Schwark et al. (2009-this issue) develop a depositional model for attenuation of the NE trade wind system over central Africa and a
the Eocene “Pesciara di Bolca” Konservat-Lagerstätte based on weakening of local upwelling conditions off NW Africa. The authors
sedimentological, palaeocological, and detailed organic geochemical invoke higher continental runoff that may have supplied excess
data. Their data indicate that the sequence was deposited in the nutrients to the eastern North Atlantic that in turn fostered greater
stagnant bottom waters of a lagoonal-like basin located on an extended production of organic matter and finally led to black shale deposition.
carbonate platform that was sheltered from open marine waters by a Mutterlose et al. (2009-this issue) analyze the macrofaunal content
submarine threshold. The high abundance of highly branched and bulk geochemistry (δ13C, δ18O) of the expanded organic-carbon-
isoprenoids in extractable bitumens suggests that run-off from nearby rich mudstone sequence of Barremian—early Aptian age exposed in
land areas provided nutrients to support an algal community northern Germany, including the mid early Aptian “Fischschiefer”
dominated by diatoms. Plant macrofossils, amber, spores and pollen, equivalent of the OAE1a. They find that the sequence of finely
and also lipid compositions indicate notable input of land plant organic laminated black shales (“Blätterton”) with organic carbon contents as
matter to the lagoon. The redox regime in general was strongly high as 7% is interbedded with dark and carbonate-poor clays on a
reducing, as evidenced by the high concentration of sulfur vs. organic decimeter to several meter scale that record repetitive fluctuations of
carbon and excellent kerogen preservation. Molecular indices suggest salinity driven stratification of the water column in relation to more
a highly stratified water column with anoxic saline bottom water and humid conditions in the hinterland. A distinctive positive δ13C
fresher surface waters, and traces of derivatives of isorenieratane, the excursion of ~4‰ preserved in belemnite guards is found to correlate
green sulfur bacteria molecular marker, indicate euxinic conditions with the global positive δ13C excursion of mid-early Aptian age
existed at times in the photic zone. The authors conclude that the following the OAE1a, providing the first record of the positive carbon
depositional setting was comparable to the Solnhofen limestones, but excursion for the Boreal Realm. A concomitant shift in δ18O of about
without the high salinities postulated for the latter. 2‰ is interpreted as a distinct SST decrease of ~ 8 °C, a trend that is also
Moving into the Mesozoic, März et al. (2009-this issue) investigate reflected by changes in lithology and organic carbon burial from the
black shales from the Coniacian-Santonian OAE3 at two locations in thick early Barremian Hauptblätterton to thin laminites in the late
different palaeo-water depths on the Demerara Rise off Surinam to Barremian. The observations from this study suggest a direct
identify systematic variations in marine and detrital sediment contribu- temperature control on peak anoxic conditions followed by repetitive
tion, depositional processes, and bottom water redox. Using a wide and short termed redox fluctuations.
range of redox proxies, including Fe/S, P/Al, C/P, and redox-sensitive Mailliot et al. (2009-this issue) present records of calcareous
trace metals, and P speciation, electron microscopy, X-ray diffraction, nannofossil and benthic foraminifer assemblages and geochemistry of
they conclude that anoxic to sulfidic bottom water and sediment two sections from the partly enclosed Causses Basin of the epiconti-
conditions existed throughout the deposition of the black shales. These nental shelf of northwest Tethys to infer a progressive environmental
extreme redox conditions were periodically punctuated by short- deterioration passing from the Late Pliensbachian to the Early Toarcian.
termed periods with less reducing bottom waters irrespective of The data show that the palaeoenvironmental deterioration culminated
palaeo-water depth. Due to strong similarities of the studied sections during the Early Toarcian OAE with a drastic decrease in nannoplank-
with the stratigraphically older OAE2 black shale on the Demerara Rise, ton production and a temporary disappearance of benthic foramini-
the authors suggest that the primary mechanisms controlling con- fers, the latter likely in response to bottom water anoxia.
tinental supply and ocean redox were time-invariant and kept the The Palaeozoic unit starts with a contribution from Bond and Wignall
western equatorial Atlantic margin widely anoxic over millions of years. (2008, 2009-this issue), who globally revisit the Frasnian-Famennian (Late
Hetzel et al. (2009-this issue) reconstruct the dynamics of deep Devonian) mass extinction, comparing new sections in the USA and Europe
ocean redox palaeoconditions during deposition of Cenomanian- (France, Germany, Poland) with published data from locations in Canada,
Turonian OAE2 black shales on the Demerara Rise. Focusing on sulfur– Australia and China. Several high-frequency relative sea-level changes are
carbon–metal relationships and especially distribution patterns of iron common to the multiple sites, supporting a transgression-anoxia-extinction
and sulfur speciation, sulfur isotope partitioning, and enrichments of link for the late Frasnian to earliest Famennian interval.
redox-sensitive and sulfide forming trace metals, they identify euxinic Two papers deal with Upper Ordovician-Silurian sediments from
conditions with at least temporarily free hydrogen sulfide in the water the North African margin of Gondwana. Armstrong et al. (2009-this
column, similar to the modern deep Black Sea. A change to reducing but issue) examine Upper Ordovician-Silurian black shales of Jordan. They
non-sulfidic conditions to allow reductive Fe and Co mobilization in provide evidence for a consistent increase in photic zone primary
oxygen-depleted nearshore sediments is inferred from elevated Fe/Al productivity during ice melting and conclude that increased flux of
and Co/Al ratios. Extremely low Mn/Al ratios further support the sinking organic matter led to euxinia extending from the photic zone
existence of an extended coastal upwelling oxygen-minimum-zone off to the sediment water interface as evidenced by the widespread
tropical South America. Finally, the authors discuss a change in the trace presence of isorenieratane derivatives, which resulted in improved
metal inventory of seawater. They postulate the enlargement of euxinic organic matter preservation. Vecoli et al. (2009-this issue) integrate a
depositional areas at the global onset of OAE2 that would have led to a palynological and palynofacies analysis with organic carbon isotope
Preface 217

measurements in southern Tunisia. The earliest Wenlock (“Ireviken Bond, D.P.G., Wignall, P.B., 2009. Extended Abstract of “The role of sea-level change and
marine anoxia in the Frasnian-Famennian (Late Devonian) mass extinction”.
Event”) and late Ludlow (“Lau Event”) isotopic excursions are Palaeogeography, Palaeoclimatology, Palaeoecology 273, 365–367 (this issue).
documented for the first time in high-latitude Gondwana, corroborat- Brumsack, H.-J., 2006. The trace metal content of recent organic carbon-rich sediments:
ing the hypothesis that these excursions reflect global changes in the implications for Cretaceous black shale formation. Palaeogeography, Palaeoclima-
tology, Palaeoecology 232, 344–361.
oceanic system. Both isotopic excursions are interpreted within a wide Capozzi, R., Negri, A., 2009. Role of sea-level forced sedimentary processes
palaeogeographic scenario. The authors propose an extended period on the distribution of organic carbon-rich marine sediments: a review of the Late
of black shale deposition from Rhuddanian to early Wenlock times Quaternary sapropels in the Mediterranean Sea. Palaeogeography, Palaeoclimatol-
ogy, Palaeoecology 273, 249–257 (this issue).
over the North African Gondwanan margin, associated with a coastal Challands, T.J., Armstrong, H.A., Maloney, D.P., Davies, J.R., Wilson, A., Owen, A.W., 2009.
upwelling-promoted productivity increase and a decrease in diversity Organic-carbon deposition and coastal upwelling at mid-latitude during the Upper
of the marine microplanktonic communities. Ordovician (Late Katian): A case study from the Welsh Basin, UK. Palaeogeography,
Palaeoclimatology, Palaeoecology 273, 395–410 (this issue).
In the final paper on Palaeozoic palaeoceanography, Challands et al.
Gennari, G., Tamburini, F., Ariztegui, D., Hajdas, I., Spezzaferri, S. 2009. Geochemical
(2009-this issue) discuss the origin of grey-black shale cycles within evidence for high-resolution variations during deposition of the Halocene S1
the Upper Ordovician (Upper Katian) succession of the Welsh Basin. sapropel on the Cretan Ridge, Eastern Mediterranean. Palaeogeography, Palaeocli-
Using a multi-proxy approach, they document high photic zone matology, Palaeoecology 273, 239–248 (this issue).
Hetzel, A., Böttcher, M.E., Wortmann, U.G., Brumsack, H.-J. 2009. Paleo-redox conditions
productivity comparable to that of modern coastal upwelling systems during OAE 2 reflected in Demerara Rise sediment geochemistry (ODP Leg 207).
and persistent dysoxic to anoxic conditions in the water column, Palaeogeography, Palaeoclimatology, Palaeoecology 273, 302–328 (this issue).
interpreting these deposits to be the result of a complex interaction of Hofmann, P., Stüsser, I., Wagner, T., Schouten, S., Sinninghe Damsté, J.S., 2008. Climate–
ocean coupling off North-West Africa during the Lower Albian: the Oceanic Anoxic
productivity and preservation. Event 1b. Palaeogeography, Palaeoclimatology, Palaeoecology 262, 157–165.
Hofmann, P., Stüsser, I., Wagner, T., Schouten, S., Sinninghe Damsté, J.S. 2009. Abstract of
Acknowledgements “Climate–ocean coupling off North-West Africa during the Lower Albian: the
Oceanic Anoxic Event 1b”. Palaeogeography, Palaeoclimatology, Palaeoecology 273,
329 (this issue).
Special thanks are due to Petr Štorch, who started with us on this Kim, J.-H., Buscail, R., Bourrin, F., Palanques, A., Sinninghe Damsté, J.S., Bonnin, J.,
adventure but had to drop out because of other demands on his time. Schouten, S. 2009. Transport and depositional process of soil organic matter during
wet and dry storms on the Têt inner shelf (NW Mediterranean). Palaeogeography,
His contributions were important to the success of the EGU session.
Palaeoclimatology, Palaeoecology 273, 228–238 (this issue).
We are particularly indebted to those important individuals who Mailliot, S., Mattioli, E., Bartolini, A., Baudin, F., Pittet, B., Guex, J. 2009. Late Pliensbachian –
carefully and conscientiously reviewed the manuscripts that we Early Toarcian (Early Jurassic) environmental changes in an epicontinental basin of
NW Europe (Causses area, central France): a micropaleontological and geochemical
received for this Special Issue. We acknowledge their excellent
approach. Palaeogeography, Palaeoclimatology, Palaeoecology 273, 346–364 (this
contributions by naming them here: issue).
Alessandro Amorosi, Achim Bechtel, Carlton Brett, Hans-Jürgen März, C., Beckmann, B., Franke, C., Vogt, C., Wagner, T., Kasten, S. 2009. Geochemical
Brumsack, James Casford, Anna M. Cruse, Olaf Dellwig, Elisabetta Erba, environment of the Coniacian-Santonian western tropical Atlantic at Demerara Rise.
Palaeogeography, Palaeoclimatology, Palaeoecology 273, 286–301 (this issue).
Karl Föllmi, Oliver Friedrich, David Gallego-Torres, Luca Giusberti, Meyers, P.A., 2006. Paleoceanographic and paleoclimatic similarities between Mediterra-
Michael Hermoso, Carmen Huguet, Steve Kershaw, Christopher nean sapropels and Cretaceous black shales. Palaeogeography, Palaeoclimatology,
Junium, Timothy W. Lyons, Axel Munnecke, Alex Page, Florentin Palaeoecology 235, 305–320.
Meyers, P.A., Negri, A., 2003. Paleoclimatic and paleoceanographic records in Mediterranean
Paris, Tatsuhiko Sakamoto, Lorenz Schwark, Jan Schwarzbauer, Caro- Sapropels and Mesozoic black shales. Special Issue, Palaeogeography, Palaeoclimatology,
line Slomp, Federico Spagnoli, Helen M. Talbot, Ellen Thomas, Steven Palaeoecology 190, 1–480.
Turgeon, Paul Wignall, Jan Zalasiewicz. Morigi, C. 2009. Benthic environmental changes in the Eastern Mediterranean Sea
during sapropel S5 deposition. Palaeogeography, Palaeoclimatology, Palaeoecology
Thanks are due to the authors of the individual research papers for 273, 258–271 (this issue).
their hard work and for their efforts to meet the generally stringent Mutterlose, J., Pauly, S., Steuber, T. 2009. Temperature controlled deposition of early
time tables we imposed on them. Cretaceous (Barremian—early Aptian) black shales in an epicontinental sea.
Palaeogeography, Palaeoclimatology, Palaeoecology 273, 330–345 (this issue).
Finally we thank Femke Wallien, who encouraged us to organize a Negri, A., Wagner, T., Meyers, P.A., 2006. Causes and consequences of marine organic
third special issue dealing with “black shales”, Jules Snelleman, and carbon burial through time. Palaeogeography, Palaeoclimatology, Palaeoecology
Fred Kop, who together provided us guidance throughout our period 235, 1–320.
Schwark, L., Ferretti, A., Papazzoni, C.A., Trevisani, E. 2009. Organic geochemistry and
as guest editors.
paleoenvironment of the Early Eocene “Pesciara di Bolca” Konservat-Lagerstätte, Italy.
Palaeogeography, Palaeoclimatology, Palaeoecology 273, 272–285 (this issue).
References Stramma, L., Johnson, G.C., Sprintall, J., Mohrholz, V., 2008. Expanding oxygen-minimum
zones in the tropical oceans. Science 320 (5876), 655–658.
Vecoli, M., Riboulleau, A., Versteegh, G., Trevisani, E. 2009. Palynology, organic
Armstrong, H.A., Abbott, G.D., Turner, B.R., Makhlouf, I.M., Aminu Bayawa, M.,
geochemistry and carbon isotope analysis of a latest Ordovician through Silurian
Pedentchouk, N., Peters, H., 2009. Black shale deposition in an Upper Ordovician-
clastic succession from borehole Tt1, Ghadamis Basin, southern Tunisia, North Africa:
Silurian permanently stratified, peri-glacial basin, southern Jordan. Palaeogeogra-
palaeoenvironmental interpretation. Palaeogeography, Palaeoclimatology, Palaeoe-
phy, Palaeoclimatology, Palaeoecology 273, 368–377 (this issue).
cology 273, 378–394 (this issue).
Bond, D.P.G., Wignall, P.B., 2008. The role of sea-level change and marine anoxia in the
Frasnian-Famennian (Late Devonian) mass extinction. Palaeogeography. Palaeoclima-
tology, Palaeoecology 263, 107–118.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Preface

Phanerozoic organic-carbon-rich marine sediments: Overview and future


research challenges
A. Negri a,⁎, A. Ferretti b, T. Wagner c, P.A. Meyers d
a
Dipartimento di Scienze del Mare, Università Politecnica delle Marche, Via Brecce Bianche, 60131 Ancona, Italy
b
Dipartimento di Scienze della Terra, Università degli Studi di Modena e Reggio Emilia, Largo S. Eufemia, 41100 Modena, Italy
c
Department of Civil Engineering and Geosciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
d
Department of Geological Sciences, The University of Michigan, Ann Arbor, Michigan, 48109-1005, USA

a r t i c l e i n f o

Article history:
Received 3 October 2008
Accepted 9 October 2008

Keywords:
Anoxic oceans
Organic-carbon-rich marine sediments
Sapropel
Black shale
Cenozoic
Mesozoic
Palaeozoic

“One of the major obsessions of many early workers, to the mid- these near-modern analogs of ancient black shales were deposited
1900s, was the application of uniformitarian principles to deposi- provides a good foundation for understanding how the older sequences
tional models for black shales“ (Arthur and Sageman, 1994). may have evolved. In contrast, because far less is generally known about
the black shales of the Palaeozoic than comparable OC-rich sequences of
The purpose of this overview of organic-carbon(OC)-rich marine either the Cenozoic or the Mesozoic, a more comprehensive summary
sediments is to provide a brief but current summary of the historical and comparison of the more ancient sequences is particularly appro-
developments, principle concepts, and remaining challenges in inte- priate to the theme of this Special Issue.
grated sapropel and black shale research. As such, it provides a
substantive introduction to the Special Issue of Palaeogeography, 1. Introduction
Palaeoclimatology, Palaeoecology on “Organic Carbon Rich Sediments
through the Phanerozoic: Processes, Progress, and Perspectives”. Given Sediments rich in organic carbon are restricted to small areas along
this focused scope, the overview does not aim to be comprehensive or continental margins in the modern ocean and have rarely accumulated
complete but to provide a solid setting for the fourteen individual during the Cenozoic, yet organic-rich sediments were widely deposited
research papers that constitute this Special Issue and two previous during multiple intervals of Mesozoic–Palaeozoic time and even earlier
special issues (Meyers and Negri, 2003; Negri et al., 2006) that cover during the Proterozoic–Archean. Global marine deposits document that
research aspects complementary to this one. Like the individual episodes of accumulation of OC-rich sediments occurred in different
contributions, this introduction and overview is organized into regions and at different times. These episodes were linked to climatic
Cenozoic, Mesozoic, and Palaeozoic units. The Cenozoic and Palaeozoic and palaeoceanographic perturbations that resulted in massive fluctua-
units are deliberately larger than the Mesozoic unit, acknowledging that tions in hydrologic and nutrient cycles and in ocean chemistry and that
much ground on Mesozoic black shale is already covered in the two recurred throughout geologic time. However, the paucity of surviving
previous special issues. Because the Late Cenozoic sapropels of the Palaeozoic and earlier black shale sections makes it difficult to
Mediterranean Basin have been extensively studied, understanding how impossible to recognize the internal structure of global events that are
common in younger OC-rich sedimentary sequences.
As underlined below, the marine sections in many localities consist
⁎ Corresponding author. of intercalated carbonates and shales that may grade into pure
E-mail address: a.negri@univpm.it (A. Negri). carbonate or black shale sedimentation, often reflecting shallowing/

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.10.002
A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227 219

deepening of the depositional environments due to sea-level fluctua- suggested that differing faunal and lithological characteristics within
tions superimposed on longer term orbital climate variability. sapropelic layers of different ages indicate different mechanisms of
Evidence for these past changes occurs at different time scales and in sapropel formation. Brongersma-Sanders (1957) had postulated that
shallow to deep marine settings across all latitudes. Their temporal and bituminous rocks may be formed when “either the supply of oxygen to
spatial differences document regional overprinting and amplification the lower water layers is excessively low (persistent stagnation) or the
or attenuation of global processes. supply of dead plankton and other oxidizable material is extremely high
In the Late Cenozoic, laminated sections in expanded OC-rich (hypertrophy)”. Sigl et al. (1978) consequently inferred that the dual
sequences are sometimes found; they provide down to annual-scale mechanisms of greater production and better preservation of organic
palaeoclimate histories. Similar expressions of climate and sea-level matter may help to explain the presence of sapropelic sediments in
variability are observed, although with less conclusive time control, “preglacial” lower Pleistocene and upper Pliocene sediments recovered
in the Mesozoic, when “Oceanic Anoxic Events” (OAEs) were by DSDP Leg 13 (Ryan et al., 1973; Cita, 1973), as well as in Miocene
deposited in mid-Cretaceous marine environments that experienced sediments recovered by DSDP Leg 42. Moreover, the discovery of
a series of relatively short periods of rapid change with massive sapropelic sediments in the western Mediterranean (Kidd et al., 1978)
effects on the global carbon cycle. These short intervals evidently revealed that their deposition was not entirely restricted to the eastern
were associated with anoxic or even sulfidic (euxinic) conditions in basin as had been believed until then. Therefore, new explanations for
the water column and were accompanied by widespread extinction sapropel formation had to be considered.
events. The same mechanisms, episodes of widespread bottom water Among the various processes that have been invoked to explain
anoxia, have been postulated for Palaeozoic OC-rich sediments that sapropel formation (see Rohling, 1994, and Rohling et al., in press, for
record the long duration of these conditions and that accompany the reviews), variations on the classical “stagnation/anoxia” and the
Devonian and Permian extinction events, two of the major crises in “increased productivity” models remain the most common. According
life history. to the stagnation/anoxia model, anoxic bottom conditions are caused
After more than half a century from the first report of sapropels in by a strong stratification of the water column that prevents vertical
deep sea sediments (Kullenberg, 1952) and more than 30 years after mixing and oxygen supply to the bottom waters. In the Mediterra-
the basic paper of Schlanger and Jenkyns (1976) defining the OAEs as nean, the origin of this stratification was explained as owing to
global phenomena, the significance of the burial of large amounts of increased Nile river runoff linked to the periodic enhancement of the
organic matter in deep-sea sediments on the global carbon cycle and East African monsoons (Rossignol-Strick, 1983, 1985) and, later, by
the nature of the palaeoceanographic and palaeoclimatic conditions increased rainfall and river discharge along the northern margins of
leading to these peculiar deposits remain topics of lively debate. In this the eastern Mediterranean Sea (Cramp et al., 1988; Rohling and
overview to Phanerozoic black shales, we briefly trace the history Hilgen, 1991). In the “productivity” model, sapropel deposition is
of sapropel and black shale studies since the first discovery of these linked to enhanced organic matter flux from highly productive surface
OC-rich sequences to their modern interpretation, and we highlight waters (Calvert, 1983; Calvert et al., 1992), inasmuch as the present
recent achievements and remaining challenges in their research. production of organic matter in the eastern Mediterranean cannot
account for the high values of total organic carbon (TOC) characteriz-
2. Cenozoic organic-carbon-rich marine sequences ing these layers (Calvert, 1983). In support of this view, evidence of a
significant increase of productivity at times of sapropel deposition is
2.1. Investigations of Miocene–Holocene Mediterranean Sea sapropels revealed by palaeo-productivity proxies, such as Ba and marine barite
concentrations (e.g., Thomson et al., 1995, 1999; Martinez-Ruiz et al.,
The word “sapropel” is a combination of ancient Greek words sapros 2000, 2003; Gallego-Torres et al., 2007). To help explain the increased
and pelos, meaning “putrefaction” and “mud”, respectively. It is a term productivity, Sarmiento et al. (1988) and Howell and Thunell (1992)
used in marine geology to describe dark-coloured, fine-grained invoked a radical change from the modern anti-estuarine circulation
sediments that are rich in amorphous organic matter. The historical to an estuarine circulation in the Mediterranean Sea.
definition of Mediterranean sapropels by Kidd et al. (1978) describes After this first phase of research in which the scientific community
them as well-delimited layers within open marine sediments, with tended to assume that the various causative processes were mutually
thickness N1 cm and organic carbon content N2%. How such OC-rich exclusive, some authors (Rohling and Gieskes, 1989; Castradori, 1993;
layers were deposited in the oligotrophic Mediterranean Sea has been a Rohling, 1994; Emeis et al., 1996, 2000) proposed a mechanism resulting
subject of interest since Bradley (1938) first predicted the widespread from the combination of the two processes – stratification and
occurrence of sapropels in the Mediterranean basins from study of productivity increase – that could have been caused by an increase of
exposures in southern Italy. He interpreted the sapropels as recording nutrient input via river runoff. However, questions remain regarding the
stagnation episodes during the Quaternary. Later, Kullenberg (1952) extent and dynamics of the anoxic/dysoxic layer in the water column.
interpreted sapropel formation as resulting from interstitial salinity The low-oxygen zone has been variously described as a near-surface
variations based on cores obtained by the 1948 Swedish Deep Sea zone in which oxygen becomes depleted (Sancetta,1999; Meyers, 2006),
Expedition. A decade later, Olausson (1961) presented a glacio-eustatic a large water mass extending through much of the water column (Murat
sea-level change model for sapropel formation that was subsequently and Got, 2000; Stratford et al., 2000), and as an “anoxic blanket” that
variously reinterpreted by other authors (e.g., Chamley, 1971; Miller, exists above the sediment/water interface (Casford et al., 2003).
1972; Ryan, 1972; Nesteroff, 1973; Cita et al., 1973; McCoy, 1974). Another fundamental question has centered on what organisms
However, Olausson's classic glacio-eustatic model gradually became were responsible for the increased primary productivity and how they
insufficient to explain all aspects of sapropel formation as more evolved through time. Are all of the multiple sapropel layers derived
information about their properties accumulated. from organic matter made by the same group of primary producers? If
DSDP Leg 42 recovered sapropel layers in cores of Pliocene to not, how were environmental conditions different from location to
Quaternary sediments at multiple sites in the eastern Mediterranean. location and through time? Kemp et al. (1999), for example, propose
Some of these layers predate the onset of the strong glacial–interglacial that diatom mats formed during summer seasons of prolonged
cycles that affect eustatic sea level. Sigl et al. (1978) inferred that since stratification resulting from Nile run-off. These mats sank rapidly at
the sapropels are interbedded as dark layers in more or less normal the beginning of autumn–winter wind mixing and delivered massive
light-colored “open marine” sediments, the formation of these layers loads of organic material to bottom sediments that quantitatively
records short-lived but catastrophic alterations in oceanographic consumed the available oxygen in the water column. From mass-
conditions that were probably caused by climate changes. They further balance calculations, Kemp et al. (1999) argue that all of the organic
220 A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227

carbon retained in the sapropels could have been supplied by these down oxidation after their deposition such that they rarely maintain their
diatoms. Although the mat model seems limited only to those original thicknesses and compositional features (e.g., Löwemark et al.,
uncommon sapropels containing diatom frustules, diatom opal is 2006; Larrasoaña et al., 2006; Thomson et al., 2006). One implication of
highly soluble, and therefore sapropels lacking diatom microfossils this phenomenon is that the original thickness and TOC concentration of
may once have contained them. Following this study, Sancetta (1999) a specific sapropel layer can influence the rate and degree of post
argued that the diatom mechanism proposed might also apply to depositional oxidation. Another is that the sub-surface burial depth and
other OC-rich laminated strata, such as the 125–85 million-year-old the position of the depositional setting with respect to deep water
mid-Cretaceous black shales. circulation can both play major roles in the preservation of sapropels.
A further multi-process interpretative step is typified by Meyers Consequently, recognition of this phenomenon cautions against over-
(2006), who suggested that increased continental runoff would have interpretation of sapropel presence or absence and over-dependence on
delivered abundant nutrients that would have first stimulated algal them as geochronological tools. For example, post-depositional erasure of
productivity, magnified export of organic matter, and increased mid- a weakly developed sapropel layer could badly skew an orbitally based
water oxygen demand. The combination of surface water dilution, timescale that does not take this possibility into consideration.
which increased salinity stratification of the upper water column and A particularly interesting and important finding of OC-rich layers in
thereby discouraged mixing and mid-water ventilation, and magni- the western Mediterranean illustrates another potential problem with
fied oxygen draw-down would have intensified and expanded the over-dependence on sapropels as geochronological markers and at the
oxygen minimum zone (OMZ) such that anoxia intruded into the same time illuminates how OC-rich sediments can be deposited.
photic zone. After photosynthetic nitrogen-fixing bacteria and Rogerson et al. (2008) describe OC-rich layers in post-glacial sediments
chemosynthetic archaea became established, their primary produc- of the Alboran Basin that could be confused with the widely expressed
tivity would have first augmented and then potentially would have Holocene S1 sapropel. However, they demonstrate that the onset of the
superseded that of the algae (e.g., Kuypers et al., 2001, 2002a,b). OC-rich layers predates the S1 sapropel by some 4–5 ky. They conclude
Accordingly the shift to microbe-amplified productivity would have that deposition of the Alboran Sea OC-rich layers was the result of a strong
persisted until climate reverted to less wet conditions that forced a reduction in surface water density and a shoaling of the interface between
return to less stratified, less productive and subsequently more the intermediate and deep waters during the latest deglacial period.
oxygenated conditions in the surface ocean. Hence, the palaeoceanographic changes associated with OC-rich layers
Recent approaches to investigate and to explain how sapropels were much like those that are believed to have accompanied deposition of
were formed have evolved towards modeling methods. Studies on most of the sapropels in the greater Mediterranean Sea, but they were
thermohaline circulation suggest that a weakening of the present-day related to local processes in the Alboran Sea and not to precessional cycles.
anti-estuarine circulation in the Mediterranean can lead to the
deposition of enough organic carbon to account for the formation of 2.2. Other Cenozoic OC-rich sediments
the upper Pleistocene–Holocene sapropels S5 to S1 (Myers et al.,
2000; Stratford et al., 2000). Earlier sapropels may have similarly been Although the Mediterranean sapropels are the most studied OC-rich
deposited. More recently, Bianchi and co-authors (2006) suggested Cenozoic sequences that are associated with low-oxygen depositional
that modified thermohaline circulation supplying oxygen only to the settings, similar sequences have been deposited in other semi-enclosed
upper 500 m of the water column, when coupled with increased basins during this Era. Some examples are the Holocene Black Sea (Aksu
productivity in the photic zone, can cause the development of an et al., 1995; Giunta et al., 2007), the early Holocene Marmara Sea (Kirci-
anoxic blanket at the sea-floor. Finally Meijer and Tuenter (2007) Elmas et al., 2008), and the late-glacial Japan Sea (Tada et al., 1992; Khim
pointed out that the effects of (1) increased discharge of northern et al., 2007). In addition, modern organic carbon-rich sediments are
borderland rivers and (2) an increase in net precipitation over the sea being deposited in unusual low-oxygen environments such as Medi-
itself are of equal or even greater importance than an increase in Nile terranean anoxic brine-filled basins (cf. Cita, 1991; Negri, 1996), and in
discharge in diminishing ventilation of deep waters. silled near-shore basins such as the Santa Barbara Basin and the Cariaco
Deep-sea and outcrop sequences reveal that sapropels occur Basin. Some of these locations are considered as modern or near-
cyclically, strongly suggesting that they are related to the “Milankovitch modern analogs of the older depositional settings that favoured
cycles” linked to the Earth's orbital cycles of eccentricity (100–400 ky), improved preservation of organic matter and led to accumulations of
obliquity (41 ky) and precession (19–22 ky). Correlation of the cyclic sapropels and black shales.
occurrence of sapropels in the Mediterranean to the orbital variations OC-rich sediments have also accumulated in Cenozoic settings that
was achieved by Hilgen (1991a,b) and later refined by Hilgen et al. have not been low in oxygen. The best examples of such places are the
(1995), who developed a continuous record of sapropel occurrences as large upwelling systems on the eastern margins of the modern-day
far back as the Late Miocene and correlated each sapropel to a well Atlantic and Pacific oceans, where high rates of marine productivity are
defined precession cycle. Because these orbital variations determine the sustained by strong vertical mixing of the surface ocean. The high
global, seasonal and latitudinal distribution of solar insolation, they production of organic matter associated with eastern margin upwelling
result in repetitive climatic fluctuations. As shown by Hilgen et al. in the Atlantic started ca 7 Mya in the Late Miocene (Diester-Haass et al.,
(1995), sapropel deposition occurred during precession minima, when 2002). An earlier onset of high productivity along the eastern margin of
the northern hemisphere received maximum summer insolation and the Pacific is recorded by the Monterey Formation, which consists of
climate in the Mediterranean region was relatively wet. laminated sediment enriched in organic carbon that was deposited
The correlation of sapropel occurrences to orbital frequencies has during the Serravallian–Tortonian (14–7 Mya) Monterey event
great value to chronology because cyclo-stratigraphy in combination (Vincent and Berger, 1985). Subsequent study of the Monterey
with other information, such as biostratigraphy and other sedimento- Formation has revealed that preservation of organic carbon is linked
logical features, allows accurate age determinations, better estimates of both to orbital-scale climate changes and to regional conditions during
sedimentation rates of individual sapropels, and comparison of same- the Middle to Late Miocene (John et al., 2002). Other examples of studies
age sapropel layers in different settings. These features are ultimately into how OC-rich sediments accumulate under areas of high productivity
based on the orbital occurrence and the isochrony of each layer. include the Late Quaternary off northern Mexico (Ganeshram et al.,1999),
Recognition that the sapropel record is sensitive to post depositional the California Margin (last 60 ky; Dean, 2007). Finally, OC-rich sediments
oxidation has been growing. In recent times, a number of investigators are known to accumulate in the Norwegian Greenland Sea, even if here
have pointed out the potential of in situ oxidation in biasing the record of organic matter has a residual origin being mainly derived from fossil
sapropels. Sapropels can evidently experience variable amounts of top- reworked organic matter (last 60 ky; Wagner and Hölemann, 1995).
A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227 221

Table 1 partly emerged from the original concepts. Taking particular advan-
A summary of the mid-Cretaceous Oceanic Anoxic Events (OAEs). Approximate tage of the spectacular analytical progress in organic and isotope
absolute-age durations are from Ogg et al. (2004) and Wagner et al. (2004)
geochemistry and computer simulations over the last decade, the new
Event Common name Geologic time Absolute duration challenges are beyond traditional discussions of OAEs. They include
OAE 3 (none) Coniacian–Santonian 87.3–84.6 Ma (∼2.7 My) but are not limited to the short term dynamics of the climate system in
OAE 2 Bonarelli Event Latest Cenomanian 93.8–93.5 Ma (∼300 ky) a generally warmer world, processes and phase relationships con-
OAE 1d Breistroffer Event Late Albian 100.6–100.2 Ma (∼400 ky
necting the atmosphere with the surface ocean (Dumitrescu et al.,
OAE 1c Tollebuc Event Late Albian 103.7–103.4 Ma (∼300 ky)
2006; Forster et al., 2007; Wagner et al., 2008), orbital forcings
OAE 1b series of sub-OAEs (Beckmann et al., 2005; Gale et al., 2005; Sageman et al., 2006;
Urbino Event Early Albian 110.9–110.6 Ma (∼300 ky) Mitchell et al., 2008), regional effects on global processes, in particular
Paquier Event Early Albian 112.0–111.6 Ma (∼400 ky) the role of land–ocean interactions in driving ocean redox and shallow
Jacob Event Late Aptian 113.6–113.2 Ma (∼400 ky)
OAE 1a Selli Event Early Aptian 124.2–123.4 Ma (∼800 ky)
ocean circulation (Nederbragt et al., 2004; Wagner et al., 2004; Mort
et al., 2007), nano-scale organo-mineral relationships (Kennedy et al.,
2002), redox variability and its impact on element and nutrient
cycling (Crusius et al., 1996; Pearce et al., 2008; Jenkyns et al., 2007;
Episodes of regionally confined accumulations of OC-rich sediments März et al., 2008), the role of microbial life in sustaining conditions
also occurred earlier in the Cenozoic. Intervals of “black shales” have favouring formation of black shale and utilizing carbon from them
been recovered in Arctic sediments (Stein, 2007), the Tethyian margin in under modern conditions (deep biosphere) (Krumholz et al., 1997;
the North between the Crimea and Uzbekistan (Gavrilov et al., 1997) and Coolen et al., 2002; D'Hondt et al., 2004; Arndt et al., 2007), triggers,
the South in Egypt and Israel (Speijer and Wagner, 2002) that are related thresholds and time scales for punctuated climate perturbations
to the 6 My Paleocene–Eocene Thermal Maximum period of global involving the emplacement of large igneous provinces (Sen et al.,
warming (Zachos et al., 2008). Off equatorial West Africa porcellanites 2006, Kuroda et al., 2007; Turgeon and Creaser, 2008)and biotic
with enhanced organic carbon concentrations and good hydrocarbon responses (Larson and Erba, 1999; Leckie et al., 2002), methane
potential were deposited during the Middle to Late Eocene (Wagner, outbursts (see review in Jenkyns, 2003; Kemp et al., 2005), wide-
2002). Interesting shallow water Eocene laminated sediments exist in spread but short term polar glaciation (Bornemann et al., 2008), and
Tunisia (Henchiri, 2007). Sometimes these events are not fully finally high-quality sedimentary climate records linked with biogeo-
developed (i.e., Eocene–Oligocene; Coccioni and Galeotti, 2003). chemical and global climate models (Bice et al., 2006; Flögel and
Rhythmically bedded sediments occur in the Oligocene–Miocene Wagner, 2006; Wagner et al., 2007), also allowing projections into the
sediments of the Paratethys that are postulated to be an earlier analog future (Flögel et al., 2008). In sum, these challenges address questions
of the modern Black Sea (i.e., Kruge et al., 1996; Schulz et al., 2005). of global relevance where progress is still to be made. Progress often is
hampered by the lack of basic constraints, including accurate chro-
3. Mesozoic organic-carbon-rich marine sequences nology, knowledge of sedimentary parameters, and – importantly –
the lack of continuous expanded sediment sections necessary to
Extensive sequences of marine black shales containing 2 to 30% extract sub-Milankovitch scale (centennial–millennial) climate change
organic carbon were episodically deposited from 183 My to 83 My in the past. Despite gaps, recent achievements like those presented in
(Toarcian through Santonian ages) during the Mesozoic Era. Continuing this Special Issue show various promising avenues where Cretaceous
investigations of the nature of the marine black shales and the processes black shale research in its wider sense is heading.
leading to their formation have placed black shale research into the
centre of today's debate on global warming and climate change. The 4. Palaeozoic organic carbon rich sequences
foundations for the current focus on extreme warm climate, rapid
biogeochemical cycling and associated climate/environmental change 4.1. A global view
were established half a century ago when petroleum industries jointly
developed the first integrated concepts to assess and predict general The Palaeozoic, spanning from 542 to 251 My, covers a period of
environmental boundary conditions and preferential depositional about 290 My, exceeding the combined duration of the Mesozoic and
settings for oil-prone source rocks (e.g. the “Organic Facies” concept Cenozoic eras. This long time interval has been characterized by a
summarized by Jones, 1987). In parallel the rise of isotope and organic variety of environmental and palaeogeographic settings, for which
geochemistry enabled academia to postulate the fundamental concept any generalization is often too restrictive. Nevertheless, most of the
of the “Oceanic Anoxic Events”, or OAEs (Schlanger and Jenkyns, 1976) Palaeozoic from the Cambrian to the Carboniferous, has been com-
that are listed in Table 1. Both basic concepts have not lost attraction over monly regarded as a global greenhouse period (Berner, 1994; Berner
time. Indeed, they have been merged with other critical aspects of Earth and Kothavala, 2001), interrupted by major but short-lived ice ages in
and Life Science and have substantially evolved into what is today a truly the Hirnantian (Late Ordovician), in the Late Devonian and in the Early
cross-disciplinary strategy to understand and quantify key elements of Carboniferous (Achab and Paris, 2007).
the Earth's climate system and how it interacts with the global carbon The whole Palaeozoic is punctuated by a profusion of episodes of
cycle. black shale deposition that represent a common and not unusual
Current research into the mid-Cretaceous OAEs serves as a focal sediment for that time. Sections in many localities consist of intercalated
point for research into extreme climate in a broader sense. Together carbonates and shales that may grade into pure carbonate or black shale
with detailed investigation into the Late Paleocene Thermal Maximum sedimentation, often and simply reflecting shallowing/deepening of the
(e.g., Speijer and Wagner, 2002; Zachos et al., 2005; Higgins and depositional environment due to sea-level fluctuations. Most of the
Schrag, 2006; Zachos et al., 2006; Schouten et al., 2007; Smith et al., attention that has so far been given to OC-rich deposits in the Palaeozoic
2007) and, although less well publicized, into OAE equivalents from has mostly reflected economic interests in hydrocarbons and metals.
the Jurassic (Hesselbo et al., 2000; Kemp et al., 2005; Cohen et al., Only recently a pure geochemical approach, including the application of
2007; Hesselbo et al., 2007; Pearce et al., 2008), the Miocene the three broad categories of biomarkers, carbon isotopes and elemental
(Woodruff and Savin, 1991; Raymo, 1994; Pagani et al., 1999; Föllmi compositions, has been attempted in order to understand the genesis
et al., 2005; Holbourn et al., 2007; Vincent and Berger, 1985), and the and significance of these peculiar deposits.
Plio–Pleistocene (see recent summary by Meyers, 2006), Cretaceous According to Nowak (2007), “black shales are most often described as
black shale research today addresses a wide range of questions that argillaceous, argillaceous–pelitic, argillaceous–siliceous and argillaceous–
222 A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227

carbonate sediments with higher amounts of more or less transformed The Devonian was a time of widespread formation of black shales.
organic matter responsible for the black or dark grey colour of these Among these, Late Devonian OC-rich sediments have attracted special
sediments.” The abundance of organic matter does not, per se, imply black consideration as they coincide with mass-extinction pulses and global
shales. The Palaeozoic, in fact, is also characterized by fossiliferous OC-rich signals of marine anoxia. Two distinct black OC-rich horizons (Kellwas-
limestones (e.g. the Silurian–Devonian Orthoceras limestones bordering ser horizons) characterize Late Frasnian sequences of Central Europe and
northern Gondwana), known for their black colour and whose high North Africa and have recently been documented also in the United
organic content is revealed by a peculiar bituminous smell and by the States (e.g., Bond and Wignall, 2005; de la Rue et al., 2007). Lower
common inclusion of soft hydrocarbon nuclei. Frasnian black shales, with TOC up to 14%, deposited diachronously in
Going ever deeper into the past, two factors keep playing a more several North African areas and also in Europe (Lüning et al., 2004), and
and more fundamental role: preservation and time resolution. OC rich were associated with an anoxic phase (Frasnes event) preceding the
sediments, either in form of black shales or limestones, do not Kellwasser events. In deeper basinal-settings, anoxic sediments rich in
necessarily reflect periods of elevated deposition of high organic organic matter persisted for longer periods till the earliest Famennian
matter but may paradoxically simply represent times of better organic (Joachimski et al., 2001; Bond et al., 2004). Another global event, the
matter preservation. In this perspective, at least, a simple measure of Hangenberg event, occurred in the latest Devonian in association with
the total organic content of the sample appears insufficient to define the second strong pulse of the Devonian extinction.
high deposition of organic matter. Then, even well-dated sequences Several studies have dealt with these deposits reflecting, again, the
do not offer the high-resolution records needed to fully document or usual dilemma between the two end-member factors: “preservation”
delineate short-time processes. Within these limitations, at least three versus “productivity” models. de la Rue et al. (2007) recently approached
major black-shale depositional events appear to have particularly the Frasnian/Famennian boundary with a multiproxy study (by the use of
attracted the attention of Palaeozoic specialists. chemical, sedimentological and biotic proxies) of the New Albany Shale, in
Indiana, documenting a drastic change from acritarch- to prasinophyte-
4.2. Carboniferous–Permian dominated associations across the F/F boundary, indicating basinal
deeper-water settings receiving a continuous input of in situ marine-
In the Early Carboniferous, many large continents were concen- derived organic matter in anoxic conditions. Bond et al. (2004) referred the
trated in the southern hemisphere, where an ice cap had covered the Kellwasser events, studied in a broad range of European occurrences, to
South Pole. The oceans between Euroamerica and Gondwana began to intense and widespread marine anoxia. In particular, the upper episode
close and, by the end of the period, the western half of Pangea was was interpreted as “an epicontinental seaway phenomenon, caused by the
already assembled. During the Permian, other major land masses upward expansion of anoxia from deep basinal locales rather than an
collided with the final assembly of the single Supercontinent of “oceanic” anoxic event that has spilled laterally into epicontinental
Pangea, spreading over the equator with a pole to pole distribution. A settings.” The same anoxic event appears to have been the only of the two
new ocean was rapidly growing on its southern end, the Tethys, ready Kellwasser episodes to be synchronous either across Europe (Bond et al.,
to assume a dominant role in the Mesozoic. 2004) or the western United States (Bond and Wignall, 2005).
Two well-known OC-rich sediments developed in the Late Palaeozoic Joachimski et al. (2001) analyzed geochemically the Kellwasser
of Central Europe, respectively in the Late Carboniferous–Early Permian horizons in Poland, there developed in carbonate sequences domi-
and in the Late Permian, the latter known as Kupferschiefer (TOC up to nated by cephalopod limestones (TOC up to 4.9%), owing to tropical to
20%) and famous for the richness in copper and silver. A recent subtropical palaeolatitudes (Marynowski et al., 2000). According to
petrographic study of the organic matter preserved in these levels Joachimski et al. (2001), the short term transgressive–regressive
from SW Poland (Nowak, 2007) associates the older shales to an open- pulses identified at a global scale by Johnson et al. (1985) caused the
lacustrine depositional environment with organic matter mostly of spread of anoxic water in formerly well-aerated environments (with
sapropelic origin. A peculiar distinct lamination of microscopic algal-rich episodic euxinic conditions into the photic zone) and intensified
laminae intercalated with thicker clay-rich laminae clearly indicates a continental runoff, with consequent higher supply of continent-
seasonal deposition. The Late Permian OC rich sediments, having derived nutrients and increase in primary productivity. Evidences of
bituminite as major organic component, indicate deposition in a shallow photic zone euxinia and wildfires have been recently documented in
marine depositional environment with very small terrestrial input Poland from a black shale horizon equivalent to the Hangenberg
(Nowak, 2007). global event (Marynowski and Filipiak, 2007).
Herbert and Compton (2007) recently confirmed deposition in a Buggisch et al. (2008) recently related three latest Devonian to
glacio-lacustrine environment for lower Permian sediments of the earliest Pennsylvanian positive excursions in δ13C, paralleled by the
Karoo Basin of South Africa, including black shale units and documenting deposition of black shales in central, northwest and southern Europe,
transition from full glacial to non-glacial conditions. High surface-water to concordant shifts in δ18Oapatite interpreted as signals of climatic
productivity associated with these OC rich shales had been deduced by cooling and ice accumulation at high latitudes.
Faure and Cole (1999) in all southwestern Gondwana basins. Piper and
Perkins (2004) applied a depositional model, established from trace- 4.4. Ordovician
element accumulation rates in the Cariaco Basin, to the Permian
Phosphoria Formation of northwest United States in order to unravel The Ordovician was a period of intense plate dispersion with land
the palaeoceanography of its deposition. The investigated Permian unit is masses mostly concentrated in the southern hemisphere and a vast
represented by a black phosphate deposit containing up to 15% organic oceanic area occupying the northern half of the globe. The different grade
carbon and deposited under conditions of O2 depletion. The study of of dispersion of the palaeoplates produced a diverse degree of
Piper and Perkins (2004) led to recognition that the hydrography of the endemicity among marine biota. In addition, the new land-mass
Permian Phosphoria Basin corresponds to modern upwelling environ- configuration led to the establishment of an active thermohaline oceanic
ments, with a moderate primary productivity. circulation by the Middle–Late Ordovician resulting in an effective
mixing of the pelagic elements and in an intense upwelling able to
4.3. Devonian amplify phytoplankton production (Achab and Paris, 2007). By the end of
the period, Gondwana regions were located at high latitudes in cold
In the Devonian, many of the Early Palaeozoic oceans were closing. climates during the onset of a severe glaciation with the rapid build-up of
Laurasia and Baltica collided, forming Laurussia, while the southern a huge ice-cap. Other areas occupied tropical and subtropical latitudes,
continents remained tied together in the supercontinent of Gondwana. with carbonate deposits dominating in Baltica and, partially, in Laurentia.
A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227 223

The glaciation at the end of the Ordovician has attracted interest not Phanerozoic, revealed that the rise of vascular plants in the Palaeozoic
only for its association with a major extinction event, but also because it “had one of the most dramatic effects on CO2 of any process occurring
required unusual causes due to its short duration (approximately 1 My) within the past 550 Ma.” Important drops in CO2 occurred in the
during an otherwise warm period in Earth's climate, characterized by an Devonian, due to intensive weathering of silicate rocks by deeply rooted
atmospheric CO2 concentration that was 15–20 times its present level plants, and in the Carboniferous and Permian, due to enhanced burial of
(Page et al., 2006). Several other minor glacial episodes appear to have organic matter resulting probably from production of microbially
been associated, and Page et al. (2006, 2007) recently referred to a glacial resistant plant remains such as lignin. By the Late Palaeozoic, either
Early Palaeozoic Icehouse (Late Ordovician–early Silurian), marked by marine or terrestrial organisms were able to contribute to organic-rich
seven glacial maxima, intercalated by brief intervals of glacial ameliora- sediments. At the end of the Permian, the most severe mass extinction of
tions. OC-rich black shales containing up to 10% TOC were deposited the life-history took place and a completely different life-scenario was
extensively especially along the northern Gondwana shelf during the opened.
same time interval (e.g., Lüning et al., 2000; Armstrong et al., 2005; Le In this context, which were the major producers of Palaeozoic organic
Heron et al., 2008). The short duration of the glaciation was explained by matter? Schwark and Empt (2006) investigated sterane compositions in
a rapid drawdown of atmospheric CO2 below a critical glacial inception Late Ordovician to Late Permian rock samples in order to unravel the
threshold triggered by an amplification in marine productivity and evolution and diversification of Palaeozoic phytoplankton (mainly
consequent organic carbon burial (Brenchley et al., 1994, 1995). The represented by Cyanophytes, acritarchs and algae) and to recognize
increased phytoplankton production eventually led to the replacement any signal recorded by algae that might document causes/effects of the
of greenhouse by icehouse conditions (Achab and Paris, 2007). A similar extinction events. Their data revealed a stepwise but persistent increase
approach was advanced by Page et al. (2006), who interpreted the in C28/C29-sterane ratio caused by a drastic change in the community
deposition of graptolitic black shales as a negative feedback mechanism structure of green algae at the Devonian/Carboniferous boundary and
responsible for drawing down CO2 and preventing runaway melting in a contemporary to a great decline of acritarchs. More primitive C29 sterane
peculiar climatic situation in which elevated atmospheric CO2 prohibited producing algae were in fact replaced by modern C28-sterane-producing
carbonate formation. algae. This event has the same fundamental significance as more recent
Armstrong et al. (2005, 2006) discussed and compared the floral revolutions like the appearance of haptophyte algae or diatoms
upwelling and the silled basin models (either the elevated productiv- (Schwark and Empt, 2006). A “phytoplankton blackout” resulting also
ity or the enhanced preservation model) in order to interpret the from a massive drop of acritarchs at the Devonian/Carboniferous
deposition of Late Ordovician deglacial organic-rich black shales from boundary developed between the Early Carboniferous and the Late
Jordan. They favoured the “expanding puddle model” of Wignall Triassic. A similar gap was not reported among benthic invertebrates. The
(1994), suggesting that anoxic conditions were maintained for strong and persistent increase in the C28/C29-sterane ratio signal
thousands of years due to a pronounced salinity stratification with a occurring at the Devonian/Carboniferous could reveal the appearance
surface layer of fresher water derived by ice-melting. An intensified of a non encysting, hardly fossilizing algal group, as earlier suggested by
bioproductivity was sustained by continuous supply of meltwater- Strothers (1996).
derived nutrients in the mixolimnion. With the onset of the deglacial In the marine realm, graptolites also made an important contribu-
transgression, black shales extended to marginal areas. tion. Organic matter preserved in lower Silurian black shales from the
A different view had been proposed by Lüning et al. (2000), and Barrandian (Czech Republic) consists primarily of graptolites and
discussed in Lüning et al. (2006), who favoured the basal transgressive subordinate chitinozoans (Suchý et al., 2002).
black shale model for earliest Silurian oil shales from Arabia and North The study of terrestrial organic matter has recently attracted interest
Africa, source of 80–90% Palaeozoic hydrocarbons of North Africa and as a more detailed knowledge of terrestrial organic matter allows a
with up to 17% total organic content (Boote et al., 1998). They correlation between terrestrial and oceanic realms, in order to determine
assembled a huge quantity of information, including unpublished if oceanic carbon perturbations are purely oceanographic in their extent
petroleum exploration well data, and suggested that the melting of or if they have affected the entire ocean–atmosphere system. Aromatic
the Late Ordovician ice-cap produced a very rapid transgression in a hydrocarbon biomarker analysis was performed on organic matter of
strongly irregular palaeorelief. An almost instantaneous anoxic event terrestrial origin from Middle Devonian to Late Permian coal basins of the
was induced by a favourable combination of several factors in the Euramerian flora realm (Armstroff et al., 2006), mostly consisting of a
Rhuddanian, persisting just 1–2 My and resulting in the excep- mixture of type II–III kerogen with TOC up to 81.5%. According to
tional preservation of the laterally discontinuous northern Gondwana palynological and palaeobotanical studies on Late Carboniferous Eur-
OC-rich basal Silurian shales. american coal basins (Auras et al., 2006), a major change in the
composition of coal-swamp floras happened during the Pennsylvanian,
4.5. Palaeozoic organic matter producers when the lycopod-dominated flora was replaced by a tree fern-dominated
flora, and seed plants assumed a prominent role. Carbon isotopic ratios
The Palaeozoic marks an interval of extraordinary and unparalleled indicate that a C3-photosynthetic pathway analogous to that of present C3-
achievements in the history of life. In the Early Palaeozoic, life existed plants had been already developed by the Late Carboniferous–Permian
in a solely marine scenario. By the Silurian, a dramatic innovative event (Auras et al., 2006). Organic-geochemical investigations carried on
was quickly occurring, with terrestrial environments that were subject Devonian, Carboniferous and Permian coals in a wide palaeogeographic
to the rapid colonization and diversification of land plants. From then scenario revealed that no significant changes in chemical composition of
on, land plants strongly became able to modify the global carbon cycle the organic matter accompanied the Late Palaeozoic rapid morphologic
and consequently the climatic history of the Earth, removing CO2 from evolution of higher land plants (Wollenweber et al., 2006). Furthermore,
the atmosphere and resulting in a significant and persistent storage of land plant organic matter from a variety of late Silurian to Late Permian
CO2 starting in the Devonian (Peters-Kottig et al., 2006). Even within terrestrial successions was recently investigated by Peters-Kottig et al.
terrestrial floras, significant revolutions occurred in the Palaeozoic. (2006), revealing no remarkable influences in organic carbon isotopic
Edwards (1998) recognized four major evolutionary steps in Palaeozoic composition due to climatic conditions.
land plant evolution. Ordovician and Silurian vegetation is represented Many of the most significant black shale episodes in the Palaeozoic
mainly by spores while tracheophytes (or vascular plants) began to strictly match with major crises in the history of life. Understanding
radiate in the Late Silurian–Early Devonian. Seed-plants appeared in the what drives global diversity may be used to explain processes, such as
Late Devonian and, by the Permian, gymnosperms became dominant. mass extinctions, that control diversity and turnover at a variety of
Berner (1998), by the use of GEOCARB II applied to the whole geographic and temporal scales. But what is the role/effect of organic
224 A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227

carbon rich sediments in extinction events? The Late Ordovician two- If our target objective is global in nature and not restricted to any
phase extinction events and the multi-staged Devonian events show a period of time, we will be able to test the validity of Processes in the
pronounced but short-lived increase of the C28/C29-sterane-ratios, recent as well as its application in the past, to obtain real Progress in
reflecting possibly the development of more opportunistic species. The the knowledge of OC-rich sediments, and to gain credibility for
quick recovery of the ratio would indicate that, at least for the younger delineating true Perspectives for the future. That way integrated
crisis, the Late Devonian extinctions did not produce any lasting research into OC-rich sediments from the past may provide critical
fundamental changes in the assemblage of the phytoplanktonic green information that help to improving climate predictions into the future
algae, and the dominant phytoplankton groups regain their older beyond the IPCC time horizon.
proportions. Prasinophyte algae, abundant in sediments deposited
under oxygen depleted conditions, might have quickly adapted in the References
events and produced the recorded increase in C28 steranes, in analogy Achab, A., Paris, F., 2007. The Ordovician chitinozoan biodiversification and its leading
with the behaviour of recent prasinophytes (Schwark and Empt, 2006). factors. Palaeogeography, Palaeoclimatology, Palaeoecology 245, 5–19.
Aksu, A.E., Yaşar, D., Mudie, P.J., 1995. Paleoclimatic and paleoceanographic circumstances
leading to development of sapropel layer S1 in the Aegean Sea. Palaeogeography,
5. Final remarks Palaeoclimatology, Palaeoecology 116, 71–101.
Armstroff, A., Wilkes, H., Schwarzbauer, J., Littke, R., Horsfield, B., 2006. Aromatic
The aim of this Special Issue is to describe and discuss Phanerozoic hydrocarbon biomarkers in terrestrial organic matter of Devonian to Permian age.
Palaeogeography, Palaeoclimatology, Palaeoecology 240, 253–274.
OC-rich sediments while breaking the walls that have traditionally Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A.,
existed between Palaeozoic, Mesozoic and Cenozoic specialists, who too Abu Salah, A., 2005. Origin, sequence stratigraphy and depositional environment of
often confine and organize themselves into hermetic “time capsules”. an upper Ordovician (Hirnantian) deglacial black shale, Jordan. Palaeogeography,
Palaeoclimatology, Palaeoecology 220, 273–289.
During the process of writing this overview and editing this Special Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A.,
Issue, we have realized that some substantial problems still hamper the Abu Salah, A., 2006. Reply to “Origin, sequence stratigraphy and depositional
Scientific Community in its approach to the study of OC-rich sediments. environment of an upper Ordovician (Hirnantian) deglacial black shale, Jordan”.
Palaeogeography, Palaeoclimatology, Palaeoecology 230, 356–360.
In general, specialists of each Era work in completely different time Arndt, S., Brumsack, H.-J., Wirtz, K.W., 2007. Cretaceous black shales as active
scales that range from thousands to millions of years. This generalization bioreactors: a biogeochemical model for the deep biosphere encountered during
implies that the “temporal magnification” through which scientists are ODP Leg 207 (Demerara Rise). Geochimica et Cosmochimica Acta 70, 408–425.
Arthur, M.A., Sageman, B.B., 1994. Marine black shales: a review of depositional
facing the subject varies from the fine resolution of the Modern and Late mechanisms and significance of ancient deposits. Annual Review of Earth and
Holocene to the extremely low resolution of the Early Palaeozoic. Planetary Sciences 22, 499–551.
Speaking metaphorically, it is like studying the same object by SEM, with Auras, S., Wilde, V., Hoernes, S., Scheffler, K., Püttmann, W., 2006. Biomarker composition
of higher plant macrofossils from Late Palaeozoic sediments. Palaeogeography,
an optical microscope, or with naked eyes and comparing their results.
Palaeoclimatology, Palaeoecology 240, 305–317.
In addition, too often we recognized an exasperating use of the Beckmann, B., Flögel, S., Hofmann, P., Wagner, T., 2005. Mid-Cretaceous African climate
uniformitarianism principle in which models or opinions derived from development and implications for the marine carbon cycle. Nature 437, 241–244.
recent examples are simplistically applied to any of the older “time- Berner, R.A., 1994. GEOCARB II: a revised model for atmospheric CO2 over Phanerozoic
time. American Journal of Science 294, 56–91.
boxes”. In actuality, physical and biological conditions (e.g., oxygen Berner, R.A., 1998. The carbon cycle and CO2 over Phanerozoic time: the role of
and CO2) have strongly varied through time. Palaeozoic black shales land plants. Philosophical Transactions of the Royal Society of London B 353,
were clearly deposited in a CO2-dominated setting (see Berner, 1994, 75–82.
Berner, R.A., Kothavala, Z., 2001. GEOCARB III: a revised model of atmospheric CO2 over
1998), whereas younger deposits reflect a lower concentration of the Phanerozoic time. American Journal of Science 301, 182–204.
same gas. Again, the nature of primary producers is not yet completely Bianchi, D., Zavatarelli, M., Pinardi, N., Capozzi, R., Capotondi, L., Corselli, C., Masina, S.,
defined for pre-Jurassic production of organic matter. Furthermore, 2006. Simulation of ecosystem response during the sapropel S1 deposition event.
Palaeogeography, Palaeoclimatology, Palaeoecology 235, 265–287.
palaeogeographic scenarios reveal completely different worlds in Bice, K.L., Birgel, D., Meyers, P.A., Dahl, K.A., Hinrichs, K.U., Richard, D.N., 2006. A
terms of land masses, oceans, palaeolatitudes, etc. According to this, multiple proxy and model study of Cretaceous upper ocean temperatures and
any attempt to model the deposition of OC-rich sediments through the atmospheric CO2 concentrations. Paleoceanography 21. doi10.1029/2005PA001203.
Bond, D., Wignall, P.B., 2005. Evidence for Late Devonian (Kellwasser) anoxic events in
Phanerozoic must necessarily be tuned with all these variables. the Great Basin, western United States. In: Over, J., Morrow, J., Wignall, P.B. (Eds.),
Another relevant point is that some of the Phanerozoic OC-rich Development in Palaeontology and Stratigraphy. Understanding Late Devonian and
sediments are defined as global events, like the Cretaceous OAE1a and Permian–Triassic Biotic and Climatic Events, 20. Elsevier, pp. 225–261 (Chapter 9).
Bond, D., Wignall, P.B., Racki, G., 2004. Extent and duration of marine anoxia during the
OAE2, but some others appear to have had a more restricted and even
Frasnian–Famennian (Late Devonian) mass extinction in Poland, Germany, Austria
localized significance. Studies on OC-rich sediments have so far been and France. Geological Magazine 141, 173–193.
concentrated on specific areas or precise time slices (e.g., the Cretaceous Boote, D.R.D., Clark-Lowes, D.D., Traut, M.W., 1998. Palaeozoic petroleum systems of North
and in here even more focused on specific isotopic excursions preserved Africa. In: MacGregor, D.S., Moody, R.T.J., Clark-Lowes, D.D. (Eds.), Petroleum Geology
of North Africa. Geological Society of London, Special Publication, 132, pp. 7–68.
in the sediments), while other periods of time have received much less Bornemann, A., Norris, R.D., Friedrich, O., Beckmann, B., Schouten, S., Sinninghe Damsté,
attention. What clearly emerges now is the urgent need of recognizing J.S., Vogel, J., Hofmann, P., Wagner, T., 2008. Isotopic evidence for glaciation during
which of these deposits are really local and which are really global the Cretaceous supergreenhouse. Science 319, 189–192.
Bradley, W.H., 1938. Mediterranean sediments and Pleistocene sea levels. Science 88,
(meaning that global is independent of latitude or physiography). This 376–379.
different attitude will probably require us to apply different approaches Brenchley, P.J., Carden, G.A.F., Marshall, J.D., 1995. Environmental changes associated
in search of possible interpretations and perhaps diverse mechanisms with the “first strike” of the Late Ordovician mass extinction. Modern Geology 20,
69–82.
leading to the deposition of OC-rich sequences. Brenchley, P.J., Marshall, J.D., Carden, G.A.F., Robertson, D.B.R., Long, D.G.F., Meidla, T.,
The three main issues described here need to be further Hints, L., Anderson, T.F., 1994. Bathymetric and isotopic evidence for a short-lived
investigated and are certainly worth answering. They should not be Late Ordovician glaciation in a greenhouse period. Geology 22, 295–298.
Brongersma-Sanders, M., 1957. Mass mortality in the sea. In: Hedgpeth, P.W. (Ed.),
regarded as problems but, on the contrary, as opportunities for new
Treatise on Marine Ecology and Paleoecology, Part 1. Geological Society of America
achievements, in particular in the light of the current discussion on Memoir, 67, pp. 941–1010.
global warming. In our opinion, the Scientific Community must come Buggisch, W., Joachimski, M.M., Sevastopulo, G., Morrow, J.R., 2008. Mississippian δ13Ccarb
and conodont apatite δ18O records—Their relation to Late Palaeozoic Glaciation.
to a multiple-time scale approach and to a constructive dialogue that
Palaeogeography, Palaeoclimatology, Palaeoecology 268 (3–4), 273–292. doi:10.1016/j.
better integrates data and models in order to be even more successful. palaeo.2008.03.043.
These efforts, with an emphasis on the upscaling/downscaling of Calvert, S.E., 1983. Geochemistry of Pleistocene sapropel and associated sediments from
processes and effects/feedbacks, will lead to the identification of the Eastern Mediterranean. Oceanologica Acta 6, 255–267.
Calvert, S.E., Nielsen, B., Fontugne, M.R., 1992. Evidence from nitrogen isotope ratios for
methodologies that may be used uniformly in the Cenozoic, Mesozoic enhanced productivity during formation of eastern Mediterranean sapropels.
and Palaeozoic. Nature 359, 223–225.
A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227 225

Casford, J.S.L., Rohling, E.J., Abu-Zied, R.H., Fontanier, C., Jorissen, F.J., Leng, M.J., Gavrilov, Y.O., Kodina, L.A., Lubchenko, I.Y., Muzylev, N.G., 1997. The late Paleocene anoxic
Schmiedel, G., Thomson, J., 2003. A dynamic concept for eastern Mediterranean event in epicontinental seas of peri-Tethys and formation of the sapropelite unit:
circulation and oxygenation during sapropel 400 formation. Palaeogeography, sedimentology and geochemistry. Lithology and Mineral Resources 32, 427–450.
Palaeoclimatology, Palaeoecology 190, 103–119. Giunta, S., Morigi, C., Negri, A., Guichard, F., Lericolais, G., 2007. Holocene biostrati-
Castradori, D., 1993. Calcareous nannofossils and the origin of eastern Mediterranean graphy and paleoenvironmental changes in the Black Sea based on calcareous
sapropels. Paleoceanography 8, 459–471. nannoplankton. Marine Micropaleontology 63, 91–110.
Chamley, H., 1971. Recherches sur la sedimentation argileuse en Méditerranée. Sciences Henchiri, M., 2007. Sedimentation, depositional environment and diagenesis of Eocene
Géologiques Strasbourg, Mémoire 35, 1–225. biosiliceous deposits in Gafsa basin (southern Tunisia). Journal of African Earth
Cita, M.B.,1973. Inventory of biostratigraphical findings and problems. In: Ryan, W.B.F., Hsü, Sciences 49, 187–200.
K.J., et al. (Eds.), Initial Reports of the Deep Sea Drilling Project, vol. 13, pp. 1045–1073. Herbert, C.T., Compton, J.S., 2007. Depositional environments of the lower Permian
Cita, M.B., 1991. Anoxic basins of the eastern Mediterranean: an overview. Paleoceano- Dwyka diamictite and Prince Albert shale inferred from the geochemistry of early
graphy 6, 133–141. diagenetic concretions, southwest Karoo Basin, South Africa. Sedimentary Geology
Cita, M.B., Chierici, M.A., Ciampo, G., Zei, M., D'Onofrio, S., Ryan, W.B.F., Scorziello, R., 1973. 194, 263–277.
The Quaternary record in the Thyrrhenian and Ionian basins of the Mediterranean Sea. Hesselbo, S.P., et al., 2000. Massive dissociation of gas hydrate during a Jurassic oceanic
In: Ryan, W.B.F., Hsü, K.J., et al. (Eds.), Initial Reports of the Deep Sea Drilling Project, 13, anoxic event. Nature 406, 392–395.
pp. 1263–1339. Hesselbo, S.P., Jenkyns, H.C., Duarte, L.V., Oliveira, L.C.V., 2007. Carbon-isotope record of
Coccioni, R., Galeotti, S., 2003. Deep-water benthic foraminiferal events from the the Early Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine
Massignano Eocene/Oligocene Boundary Stratotype Section and Point (central carbonate (Lusitanian Basin, Portugal). Earth and Planetary Science Letters 253,
Italy): biostratigraphic, paleoecologic, and paleoceanographic implications. In: 455–470.
Prothero, D.R., Ivany, L.C., Nesbitt, E. (Eds.), From Greenhouse to Icehouse: The Higgins, J.A., Schrag, D.P., 2006. Beyond methane: towards a theory for the Paleocene–
Marine Eocene–Oligocene Transition. Columbia University Press, pp. 438–452. Eocene Thermal Maximum. Earth and Planetary Science Letters 245, 523–537.
Cohen, A.S., Coe, A.L., Kemp, D.B., 2007. The late Paleocene–early Eocene and Toarcian (Early Hilgen, F.J., 1991a. Astronomical calibration of Gauss to Matuyama sapropels in the
Jurassic) carbon-isotope excursions: a comparison of their timescales, associated Mediterranean and implication for the geomagnetic polarity time scale. Earth and
environmental changes, causes and consequences. Journal of the Geological Society of Planetary Science Letters 104, 226–244.
London 164, 1093–1108. Hilgen, F.J., 1991b. Extension of the astronomically calibrated (polarity) to the Miocene/
Coolen, M.J.L., Cypionka, H., Sass, A.M., Sass, H., Overmann, J., 2002. Ongoing Pliocene boundary time scale. Earth and Planetary Science Letters 107, 349–368.
modification of Mediterranean Pleistocene sapropels mediated by prokaryotes. Hilgen, F.J., Krijgsman, W., Langereis, C.G., Lourens, L.J., Santarelli, A., Zachariasse, W.J.,
Science 296, 2407–2410. 1995. Extending the astronomical (polarity) time scale into the Miocene. Earth and
Cramp, A., Collins, M., West, R., 1988. Late Pleistocene–Holocene sedimentation in the Planetary Science Letters 136, 495–510.
NW Aegean Sea: a paleoclimatic–paleoceanographic reconstruction. Palaeogeo- Holbourn, A., Kuhnt, W., Schulz, M., Flores, J.A., Andersen, N., 2007. Orbitally-paced
graphy, Palaeoclimatology, Palaeoecology 68, 61–77. climate evolution during the middle Miocene “Monterey” carbon-isotope excur-
Crusius, J., Calvert, S., Pedersen, T., Sage, D., 1996. Rhenium and molybdenum sion. Earth and Planetary Science Letters 261, 534–550.
enrichments in sediments as indicators of oxic, suboxic and sulfidic conditions of Howell, M.W., Thunell, R.C., 1992. Organic carbon accumulation in Bannock Basin:
deposition. Earth and Planetary Science Letters 145, 65–78. evaluating the role of the productivity in the formation of eastern Mediterranean
Dean, W.E., 2007. Sediment geochemical records of productivity and oxygen depletion sapropels. Marine Geology 103, 461–471.
along the margin of western North America during the past 60,000 years: Jenkyns, H.C., 2003. Evidence for rapid climate change in the Mesozoic–Palaeogene
teleconnections with Greenland Ice and the Cariaco Basin. Quaternary Science greenhouse world. Philosophical Transactions of the Royal Society London, Series A 361,
Reviews 26, 98–114. 1885–1916.
de la Rue, S.R., Rowe, H.D., Rimmer, S.M., 2007. Palynological and bulk geochemical Jenkyns, H.C., Matthews, A., Tsikos, H., Erel, Y., 2007. Nitrate reduction, sulphate reduction,
constraints on the paleoceanographic conditions across the Frasnian–Famennian and sedimentary iron isotope evolution during the Cenomanian–Turonian ocean
boundary, New Albany Shale, Indiana. International Journal of Coal Geology 71, 72–84. anoxic event. Paleoceanography 22, PA3208. doi10.1029/2006PA001355.
D'Hondt, S., et al., 2004. Distributions of microbial activities in deep subseafloor Joachimski, M.M., Ostertag-Henning, C., Pancost, R.D., Strauss, H., Freeman, K.H., Littke,
sediments. Science 306, 2216–2221. R., Sinninghe Damsté, J.S., Racki, G., 2001. Water column anoxia, enhanced
Diester-Haass, L., Meyers, P.A., Vidal, L., 2002. The late Miocene onset of high productivity in productivity and concomitant changes in δ13C and δ34S across the Frasnian–
the Benguela upwelling area as part of a global pattern. Marine Geology 180, 87–103. Famennian boundary (Kowala-Holy Cross Mountains/Poland). Chemical Geology
Dumitrescu, M., Brassell, S.C., Schouten, S., Hopmans, E.C., Sinninghe Damsté, J.S., 2006. 175, 109–131.
Instability in tropical Pacific sea-surface temperatures during the early Aptian. John, C.M., Föllmi, K., De Kaenel, E., Adatte, T., Steinmann, P., Badertscher, C., 2002.
Geology 34, 833–836. Carbonaceous and phosphate-rich sediments of the Miocene Monterey Formation at
Edwards, D., 1998. Climate signals in Palaeozoic land plants. Philosophical Transactions El Capitan State Beach, California, U.S.A. Journal of Sedimentary Research 72, 252–267.
of the Royal Society of London B 353, 141–157. Johnson, J.G., Klapper, G., Sandberg, C.A., 1985. Devonian eustatic fluctuations in
Emeis, K.-C., Sakamoto, T., Wehausen, R., Brumsack, H.-J., 2000. The sapropel record of Euramerica. Geological Society of America Bulletin 96, 567–587.
the eastern Mediterranean Sea-results of Ocean Drilling Program Leg 160. Jones, R.W., 1987. Organic facies. Advances in Petroleum Geochemistry 2, 1–80.
Palaeogeography, Palaeoclimatology, Palaeoecology 158, 371–395. Kemp, A.E.S., Pearce, R.B., Koizumi, I., Pike, J., Rance, S.J., 1999. The role of mat-forming
Emeis, K.-C., Shipboard Scientific Party of Leg 160, 1996. Paleoceanography and sapropel diatoms in the formation of Mediterranean sapropels. Nature 398, 57–61.
introduction. In: Emeis, K.C., Robertson, A.H.F., Richter, C., et al. (Eds.), Proceedings Kemp, D.B., Coe, A.L., Cohen, A.S., Schwark, L., 2005. Astronomical pacing of methane
of the Ocean Drilling Program Initial Reports, vol. 160, pp. 21–27. release in the Early Jurassic period. Nature 437, 396–399.
Faure, K., Cole, D.I., 1999. Geochemical evidence for lacustrine microbial blooms in the Kennedy, M.J., Pevear, D.R., Hill, R.J., 2002. Mineral surface control of organic carbon in
vast Permian Main Karoo, Paraná, Falkland Islands and Huab basins of southwestern black shale. Science 295, 657–660.
Gondwana. Palaeogeography, Palaeoclimatology, Palaeoecology 152, 189–213. Khim, B.K., Bahk, J.J., Hyun, S., Lee, G.H., 2007. Late Pleistocene dark laminated mud layers
Flögel, S., Hofmann, P., Beckmann, B., Bornemann, A., Norris, D., Wagner, T., 2008. from the Korea Plateau, western East Sea/Japan Sea, and their paleoceanographic
Evolution of tropical watersheds and continental hydrology during the late Creta- implications. Palaeogeography, Palaeoclimatology, Palaeoecology 247, 74–87.
ceous greenhouse: marine carbon burial and possible implications for the future. Kidd, R.B., Cita, M.B., Ryan, W.B.F., 1978. Stratigraphy of eastern Mediterranean sapropel
Earth and Planetary Science Letters 274 (1–2), 1–13. doi:10.1016/j.epsl.2008.06.011. sequences recovered during Leg 42A and their paleoenvironmental significance. In:
Flögel, S., Wagner, T., 2006. Linking Mid-Cretaceous global climate modeling and Hsü, K.J., Montadert, L., et al. (Eds.), Initial Reports of the Deep-Sea Drilling Project
geological marine proxy records—insights to insolation-driven coupling of atmo- 42A. U.S. Government Printing Office, Washington, pp. 421–443.
spheric circulation and ocean properties. Palaeogeography, Palaeoclimatology, Kirci-Elmas, E., Algan, O., Özkar-Öngen, Í., Struck, U., Altenbach, A.V., Sagular, E.K., Nazíc,
Palaeoecology 235, 288–304. A., 2008. Palaeoenvironmental investigation of sapropelic sediments from the
Föllmi, K.B., Badertscher, C., de Kaenel, E., Stille, P., John, C.M., Adatte, T., Steinmann, P., Marmara Sea: a biostratigraphic approach to palaeoceanographic history during
2005. Phosphogenesis and organic-carbon preservation in the Miocene Monterey the Last Glacial–Holocene. Turkish Journal of Earth Sciences 17, 129–168.
Formation at Naples Beach, California—The Monterey hypothesis revisited. Kruge, M.A., Mastalerz, M., Solecki, A., Stankiewicz, B.A., 1996. Organic geochemistry and
Geological Society of America Bulletin 117, 589–619. petrology of oil source rocks, Carpathian Overthrust region, southeastern Poland—
Forster, A., Schouten, S., Moriya, K., Wilson, P.A., Sinninghe Damsté, J.S., 2007. Tropical implications for petroleum generation. Organic Geochemistry 24, 897–912.
warming and intermittent cooling during the Cenomanian/Turonian oceanic anoxic Krumholz, L., McKinley, J., Ulrich, G., Suflita, J., 1997. Confined subsurface microbial
event 2: Sea surface temperature records from the equatorial Atlantic. Paleoceano- communities in Cretaceous rock. Nature 386, 64–66.
graphy 22, PA1219. doi10.1029/2006PA001349. Kullenberg, B., 1952. On the salinity of the water contained in marine sediments.
Gale, A.S., Kennedy, W.J., Voigt, S., Walaszczyk, I., 2005. Stratigraphy of the Upper Göteborgs Kungliga Vetenskaps- och Vitterhets-Samhaelles Handlingar 6, 3–37.
Cenomanian–Lower Turonian Chalk succession at Eastbourne, Sussex, UK: ammonites, Kuroda, J., Ogawa, N.O., Tanimizu, M., Coffin, M.F., Tokuyama, H., Kitazato, H., Ohkouchi,
inoceramid bivalves and stable carbon isotopes. Cretaceous Research 26, 460–487. N., 2007. Contemporaneous massive subaerial volcanism and late Cretaceous
Gallego-Torres, D., Martinez-Ruiz, F., Paytan, A., Jiménez-Espejo, F.J., Ortega-Huertas, M., Oceanic Anoxic Event 2. Earth and Planetary Letters vol. 256, 211–223.
2007. Pliocene–Holocene evolution of depositional conditions in the eastern Kuypers, M.M.M., Blokker, P., Erbacher, J., Kinkel, H., Pancost, R.D., Schouten, S.,
Mediterranean: Role of anoxia vs. productivity at time of sapropel deposition. Sinninghe Damsté, J.S., 2001. Massive expansion of marine archaea during a mid-
Palaeogeography, Palaeoclimatology, Palaeoecology 246, 424–439. Cretaceous oceanic anoxic event. Science 293, 92–94.
Ganeshram, R.S., Calvert, S.E., Pedersen, T.F., Cowie, G.L., 1999. Factors controlling the Kuypers, M.M.M., Blokker, P., Hopmans, E.C., Kinkel, H., Pancost, R.D., Schouten, S.,
burial of organic carbon in laminated and bioturbated sediments off NW Mexico: Sinninghe Damsté, J.S., 2002b. Archaeal remains dominate marine organic matter
implications for hydrocarbon preservation. Geochimica et Cosmochimica Acta 63, from the early Aptian oceanic anoxic event 1b. Palaeogeography, Palaeoclimatology,
1723–1734. Palaeoecology 185, 211–234.
226 A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227

Kuypers, M.M.M., Pancost, R.D., Nijenhuis, I.A., Sinninghe Damsté, J.S., 2002a. Enhanced Ogg, J.G., Agterberg, F.P., Gradstein, F.M., 2004. The Cretaceous Period. In: Gradstein, F.,
productivity led to increased organic carbon burial in the euxinic North Atlantic Ogg, J., Smith, A. (Eds.), A Geologic Time Scale 2004. Cambridge University Press,
basin during the late Cenomanian oceanic anoxic event. Paleoceanography 17, 1051. pp. 344–383.
doi10.1029/222PA000569. Olausson, E., 1961. Studies of deep-sea cores. Report of the Swedish Deep Sea Expedition
Larrasoaña, J.C., Roberts, A.P., Hayes, A., Wehausen, R., Rohling, E.J., 2006. Detecting 1947–1948, vol. 8, pp. 353–391.
missing beats in the Mediterranean climate rhythm from magnetic identification of Pagani, M., Freeman, K.H., Arthur, M.A., 1999. Late Miocene atmospheric CO2 levels and
oxidized sapropels (Ocean Drilling Program Leg 160). Physics of the Earth and the expansion of C4 grasses. Science 285, 876–879.
Planetary Interiors 156, 283–293. Page, A., Williams, M., Zalasiewicz, J., 2006. Were transgressive black shales a negative
Larson, R.L., Erba, E., 1999. Onset of the mid-Cretaceous greenhouse in the Barremian– feedback mechanism modulating glacio-eustatic cycles in the Early Palaeozoic
Aptian: Igneous events and the biological, sedimentary, and geochemical Icehouse? Geophysical Research Abstracts 8, 09210.
responses. Paleoceanography 14, 663–678. Page, A., Zalasiewicz, J., Williams, M., 2007. Deglacial anoxia in the Early Palaeozoic
Leckie, R.M., Bralower, T.J., Cashman, R., 2002. Oceanic anoxic events and plankton Icehouse. 51st Palaeontological Association Annual Meeting, Uppsala, p. 77. 2007.
evolution: biotic response to tectonic forcing during the mid-Cretaceous. Pearce, C.R., Cohen, A.S., Coe, A.L., Burton, K.W., 2008. Molybdenum isotope evidence for
Paleoceanography 17, 1041. doi10.1029/2001PA000623. global ocean anoxia coupled with perturbations to the carbon cycle during the Early
Le Heron, D.P., Khoukhi, Y., Paris, F., Ghienne, J.-F., Le Herissé, A., 2008. Black shale, grey Jurassic. Geology 36, 231–234.
shale, fossils and glaciers: anatomy of the Upper Ordovician–Silurian succession in Peters-Kottig, W., Strauss, H., Kerp, H., 2006. The land plant δ13C record and plant
the Tazzeka Massif of eastern Morocco. Gondwana Research 14, 483–496. evolution in the Late Palaeozoic. Palaeogeography, Palaeoclimatology, Palaeoecol-
Löwemark, L., Lin, Y., Chen, H.-F., Yang, T.-N., Beier, C., Werner, F., Lee, C.-Y., Song, S.-R., ogy 240, 237–252.
Kao, S.-J., 2006. Sapropel burn-down and ichnological response to late Quaternary Piper, D.Z., Perkins, R.B., 2004. A modern vs. Permian black shale: the hydrography, primary
sapropel formation in two ∼400 ky records from the eastern Mediterranean Sea. productivity, and water-column chemistry of deposition. Chemical Geology 206,177–197.
Palaeogeography, Palaeoclimatology, Palaeoecology 239, 406–425. Raymo, M.E., 1994. The Himalayas, organic carbon burial, and climate in the Miocene.
Lüning, S., Craig, J., Loydell, D.K., Štorch, P., Fitches, B., 2000. Lowermost Silurian “hot Paleoceanography 9, 399–404.
shales” in North Africa and Arabia: regional distribution and depositional model. Rogerson, M., Cacho, I., Jimenez-Espejo, F., Reguera, M.I., Sierro, F.J., Martinez-Ruiz, F.,
Earth Science Reviews 49, 121–200. Frigola, J., Canals, M., 2008. A dynamic explanation for the origin of the western
Lüning, S., Loydell, D.K., Štorch, P., Shahin, Y., Craig, J., 2006. Origin, sequence stratigraphy and Mediterranean organic-rich layers. Geochemistry, Geophysics, and Geosystems
depositional environment of an Upper Ordovician (Hirnantian) deglacial black shale, (G3) 9. doi:10.1029/2007GC001936.
Jordan—Discussion. Palaeogeography, Palaeoclimatology, Palaeoecology 230, 352–355. Rohling, E.J., 1994. Review and new aspects concerning the formation of eastern
Lüning, S., Wendt, J., Belka, Z., Kaufmann, B., 2004. Temporal–spatial reconstruction of Mediterranean sapropels. Marine Geology 122, 1–28.
the early Frasnian (Late Devonian) anoxia in NW Africa: new field data from the Rohling, E.J., Abu-Zied, R., Casford, J.S.L., Hayes, A., Hoogakker, B.A.A., in press. The
Ahnet Basin (Algeria). Sedimentary Geology 163, 237–264. Mediterranean Sea: Present and Past. In: Woodward, J.C. (Ed.), The Physical
Martinez-Ruiz, F., Kastner, M., Paytan, A., Ortega-Huertas, M., Bernasconi, S.M., 2000. Geography of the Mediterranean, Oxford.
Geochemical evidence for enhanced productivity during S1 sapropel deposition in Rohling, E.J., Gieskes, W.W.C., 1989. Late Quaternary changes in Mediterranean
the eastern Mediterranean. Paleoceanography 15, 200–209. intermediate water density and formation rate. Paleoceanography 4, 531–545.
Martinez-Ruiz, F., Paytan, A., Kastner, M., González-Donoso, J.M., Linares, D., Bernasconi, Rohling, E.J., Hilgen, F.J., 1991. The eastern Mediterranean climate at times of sapropel
S.M., Jimenez-Espejo, F.J., 2003. A comparative study of the geochemical and formation: a review. Geologie en Mijnbouw 70, 253–264.
mineralogical characteristics of the S1 sapropel in the western and eastern Rossignol-Strick, M., 1983. African monsoons, an immediate climate response to orbital
Mediterranean. Palaeogeography, Palaeoclimatology, Palaeoecology 190, 23–37. insolation. Nature 304, 46–49.
Marynowski, L., Filipiak, P., 2007. Water column euxinia and wildfire evidence during Rossignol-Strick, M., 1985. Mediterranean Quaternary sapropels, an immediate
deposition of the Upper Famennian Hangenberg event horizon from the Holy Cross response of the African monsoon to variation of insolation. Palaeogeography,
Mountains (central Poland). Geological Magazine 144, 569–595. Palaeoclimatology, Palaeoecology 49, 237–263.
Marynowski, L., Narkiewicz, M., Grelowski, C., 2000. Biomarkers as environmental Ryan, W.B.F., 1972. Stratigraphy of Late Quaternary sediments in the Eastern
indicators in a carbonate complex, example from the Middle to Upper Devonian, Mediterranean. In: Stanley, D.J. (Ed.), The Mediterranean Sea: A Natural Sedimenta-
Holy Cross Mountains, Poland. Sedimentary Geology 137, 187–212. tion Laboratory. Dowden, Hutchinson and Ross, Stoudsburg, PA, pp. 149–170.
März, C., Poulton, S.W., Beckmann, B., Küster, K., Wagner, T., Kasten, S., 2008. Redox Ryan, W.B.F., Hsü, K.J., et al., 1973. Initial Reports of the Deep Sea Drilling Project, vol. 13.
sensitivity of P cycling during marine black shale formation: dynamics of sulfidic and U.S. Government Printing Office, Washington. 514 pp.
anoxic, non-sulfidic bottom waters. Geochimica et Cosmochimica Acta 72, 3703–3717. Sageman, B.B., Meyers, S.R., Arthur, M.A., 2006. Orbital time scale and new C-isotope
McCoy, F.W., 1974. Late Quaternary sedimentation in the eastern Mediterranean Sea. Ph.D. record for Cenomanian–Turonian boundary stratotype. Geology 34, 125–128.
Thesis, Harvard University, Cambridge, Massachusetts, 164 pp. Sancetta, C., 1999. The mystery of the sapropels. Nature 398, 27–29.
Meijer, P.Th., Tuenter, E., 2007. The effect of precession-induced changes in the Sarmiento, J.L., Herbert, T., Toggweiler, J.R., 1988. Mediterranean nutrient balance and
Mediterranean freshwater budget on circulation at shallow and intermediate episodes of anoxia. Global Biogeochemical Cycles 2, 427–444.
depth. Journal of Marine Systems 68, 349–365. Schlanger, S.O., Jenkyns, H.C., 1976. Cretaceous oceanic anoxic events: causes and
Meyers, P., 2006. Paleoceanographic and paleoclimatic similarities between Mediterra- consequences. Geologie en Mijnbouw 55, 179–184.
nean sapropels and Cretaceous black shales. Palaeogeography, Palaeoclimatology, Schouten, S., Woltering, M., Rijpstra, W.I.C., Sluijs, A., Brinkhuis, H., Sinninghe Damsté, J.S.,
Palaeoecology 235, 305–320. 2007. The Paleocene–Eocene carbon isotope excursion in higher plant organic matter:
Meyers, P.A., Negri, A. (Eds.), 2003. Paleoclimatic and Paleoceanographic Records in Differential fractionation of angiosperms and conifers in the Arctic. Earth and
Mediterranean Sapropels and Mesozoic Black Shales. Special Issue. Palaeogeogra- Planetary Science Letters 258, 581–592.
phy, Palaeoclimatology, Palaeoecology, vol. 190, pp. 1–480. Schulz, H.-M., Bechtel, A., Sachsenhofer, R.F., 2005. The birth of the Paratethys during
Miller, A.R., 1972. Speculations concerning bottom circulation in the Mediterranean Sea. the Early Oligocene: from Tethys to an ancient Black Sea analogue? Global and
In: Stanley, D.J. (Ed.), The Mediterranean Sea: A Natural Sedimentation Laboratory. Planetary Change 49, 163–176.
Dowden, Hutchinson and Ross, Stoudsburg, PA, pp. 37–42. Schwark, L., Empt, P., 2006. Sterane biomarkers as indicators of Palaeozoic algal
Mitchell, R.N., Bice, D.M., Montanari, A., Cleaveland, L.C., Christianson, K.T., Coccioni, R., evolution and extinction events. Palaeogeography, Palaeoclimatology, Palaeoecol-
Hinnov, L.A., 2008. Oceanic anoxic cycles? Orbital prelude to the Bonarelli Level ogy 240, 225–236.
(OAE 2). Earth and Planetary Science Letters 267, 1–16. Sen, G., Borges, M., Marsh, B.D., 2006. A Case for short duration of Deccan Trap eruption.
Mort, H.P., Adatte, T., Föllmi, K.B., Keller, G., Steinmann, P., Matera, V., Berner, Z., Stüben, EOS, Transactions American Geophysical Union 87, 197–204.
D., 2007. Phosphorus and the roles of productivity and nutrient recycling during Sigl, W., Chamley, H., Fabricius, F., Giroud d'Argoud, G., Müller, J., 1978. Sedimentology
oceanic anoxic event 2. Geology 35, 483–486. and environmental conditions of sapropels. In: Hsü, K.J., Montadert, L., et al. (Eds.),
Murat, A., Got, H., 2000. Organic carbon variations of the eastern Mediterranean Initial Reports of the Deep-Sea Drilling Project 42A. U.S. Government Printing
Holocene sapropel: a key for understanding formation processes. Palaeogeography, Office, Washington, pp. 445–465.
Palaeoclimatology, Palaeoecology 158, 241–257. Smith, F.A., Wing, S.L., Freeman, K.H., 2007. Magnitude of the carbon isotope excursion
Myers, P.G., Haines, K., Rohling, E.J., 2000. Modelling the paleocirculation of the at the Paleocene–Eocene thermal maximum: the role of plant community change.
Mediterranean: the last glacial maximum and the Holocene with emphasis on the Earth and Planetary Science Letters, 262, 50–65.
formation of sapropel S1. Paleoceanography 13, 586–606. Speijer, R.P., Wagner, T., 2002. Transgressive black shales associated with the Late
Nederbragt, A., Thurow, J., Vonhof, H., Brumsack, H.-J., 2004. Modelling oceanic carbon Paleocene Thermal Maximum (LPTM): organic geochemical and micropaleontolo-
and phosphorus fluxes: implications for the cause of the late Cenomanian Oceanic gical evidence from the southern Tethyan margin (Egypt). In: Koeberl, C., MacLeod,
Anoxic Event (OAE2). Journal of the Geological Society London 161, 721–728. K.C. (Eds.), Catastrophic Events and Mass Extinctions: Impacts and Beyond.
Negri, A., 1996. Possible origin of laminated sediments of the anoxic Bannock Basin Geological Society of America, vol. 356, pp. 533–549.
(eastern Mediterranean). Geo-Marine Letters 16, 101–107. Stein, R., 2007. Upper Cretaceous/lower Tertiary black shales near the North Pole:
Negri, A., Wagner, T., Meyers, P.A. (Eds.), 2006. Causes and Consequences of Marine organic-carbon origin and source-rock potential. Marine and Petroleum Geology 24,
Organic Carbon Burial Through Time. Palaeogeography, Palaeoclimatology, 67–73.
Palaeoecology, vol. 235, pp. 1–320. Stratford, K., Williams, R.G., Myers, P.G., 2000. The impact of the circulation on sapropel
Nesteroff, W.D.,1973. Petrography and mineralogy of sapropels. In: Ryan, W.B.F., Hsü, K.J., et al. formation in the Eastern Mediterranean. Global Biogeochemical Cycles 14, 683–695.
(Eds.), Initial Reports of the Deep Sea Drilling Project, vol. 13. U.S. Government Printing Strothers, P.K., 1996. Acritarchs. In: Jansonius, J., McGregor, D.C. (Eds.), Palynology:
Office, Washington, pp. 713–720. Principles and Applications. American Association of Stratigraphic Palynologist
Nowak, G.J., 2007. Comparative studies of organic matter petrography of the late Foundation, pp. 81–105.
palaeozoic black shales from Southwestern Poland. International Journal of Coal Suchý, V., Sýkorová, I., Stejskal, M., Šafanda, J., Machovič, V., Novotná, M., 2002.
Geology 71, 568–585. Dispersed organic matter from Silurian shales of the Barrandian Basin, Czech
A. Negri et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 218–227 227

Republic: optical properties, chemical composition and thermal maturity. Interna- Wagner, T., Herrle, J.O., Sinninghe Damsté, J.S., Schouten, S., Stüsser, I., Hofmann, P.,
tional Journal of Coal Geology 53, 1–25. 2008. Rapid warming and salinity changes of Cretaceous surface waters in the
Tada, R.R., Koizumi, I., Cramp, A., Rahman, A., 1992. Correlation of dark and light layers, subtropical North Atlantic. Geology 36, 203–206.
and the origin of their cyclicity in the Quaternary sediments from the Japan Sea. In: Wagner, T., Hölemann, J.A., 1995. Deposition of organic matter in the Norwegian–
Pisciotto, K.A., IngleJr. Jr., J.C., von Breymann, M.T., Barron, J., et al. (Eds.), Greenland Sea during the past 2.7 million years. Quaternary Research 44, 355–366.
Proceedings of the Ocean Drilling Program, Scientific Results, 127/128, pp. 577–601. Wagner, T., Sinninghe Damsté, J., Beckmann, B., Hofmann, P., 2004. Euxinia and primary
Thomson, J., Croudace, I.W., Rothwell, R.G., 2006. A geochemical application of the production in Upper Cretaceous eastern equatorial Atlantic surface waters fostered
ITRAX scanner to a sediment core containing eastern Mediterranean sapropel units. orbital-driven formation of marine black shales. Paleoceanography 19. doi10.1029/
In: Rothwell, R.G. (Ed.), New Techniques in Sediment Core Analysis, 267. Geological 2003PA000898.
Society of London, Special Publication, pp. 65–77. Wagner, T., Wallmann, K., Stüsser, I., Herrle, J.O., Hofmann, P., 2007. Consequences of
Thomson, J., Higgs, N.C., Wilson, T.R.S., Croudace, I.W., de Lange, G.J., van Sanvoort, P.J.M., moderate ∼25,000 yr lasting emission of light CO2 into the mid-Cretaceous ocean.
1995. Redistribution and geochemical behaviour of redox sensitive elements Earth Planetary Science Letters 259, 200–211.
around S1, the most recent eastern Mediterranean sapropel. Geochimica et Wignall, P.B., 1994. Black Shales. Clarendon Press, Oxford. 127 pp.
Cosmochimica Acta 59, 3487–3501. Wollenweber, J., Schwarzbauer, J., Littke, R., Wilkes, H., Armstroff, A., Horsfield, B., 2006.
Thomson, J., Mercone, D., de Lange, G.J., van Santvoort, P.J.M., 1999. Review of recent Characterisation of non-extractable macromolecular organic matter in Palaeozoic
advances in the interpretation of eastern Mediterranean sapropel S1 from coals. Palaeogeography, Palaeoclimatology, Palaeoecology 240, 275–304.
geochemical evidence. Marine Geology 153, 77–89. Woodruff, F., Savin, S.M., 1991. Mid-Miocene isotope stratigraphy in the deep sea: high-
Turgeon, S.C., Creaser, R.A., 2008. Cretaceous oceanic anoxic event 2 triggered by a resolution correlations, paleoclimatic cycles, and sediments preservation. Paleo-
massive magmatic episode. Nature Geosciences 454, 323–326 v. ceanography 6, 755–806.
Vincent, E., Berger, W.H., 1985. Carbon dioxide and polar cooling in the Miocene: The Zachos, J.C., et al., 2005. Rapid acidification of the ocean during the Paleocene–Eocene
Monterey Hypothesis. In: Sundquist, E.T., Broecker, W.S. (Eds.), The Carbon Cycle thermal maximum. Science 308, 1611–1615.
and Atmospheric CO2: Natural Variations Archean to Present. American Geophy- Zachos, J.C., et al., 2006. Extreme warming of mid-latitude coastal ocean during the
sical Union, Washington DC, American Geophysical Union, Geophysical Monograph Paleocene–Eocene Thermal Maximum: Inferences from TEX86 and isotope data.
32, pp. 455–468. Geology 34, 737–740.
Wagner, T., 2002. Organic sedimentation in the equatorial Atlantic: evolution from Zachos, J.C., Dickens, G.R., Zeebe, R.E., 2008. An early Cenozoic perspective on
Cretaceous to early Quaternary depositional environments. Palaeogeography, greenhouse warming and carbon-cycle dynamics. Nature 451, 279–283.
Palaeoclimatology, Palaeoecology 179, 113–147.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Transport and depositional process of soil organic matter during wet and dry storms
on the Têt inner shelf (NW Mediterranean)
J.-H. Kim a,⁎, R. Buscail b, F. Bourrin b,e, A. Palanques c, J.S. Sinninghe Damsté a, J. Bonnin b,d, S. Schouten a
a
Royal Netherlands Institute for Sea Research (NIOZ), Department of Marine Organic Biogeochemistry, PO Box 59, 1790 AB Den Burg, Texel, The Netherlands
b
CEFREM-CNRS UMR 5110, Université de Perpignan, 52 avenue Paul Alduy, 66860 Perpignan Cedex, France
c
Instituto de Ciencias del Mar-C.S.I.C., Passeig Maritim de la Barceloneta, 37, 08003 Barcelona, Spain
d
EPOC-UMR 5805, Université de Bordeaux 1, Avenue des Facultés, 33405 Talence Cedex, France
e
Observatoire Océanologique de Villefranche, Laboratoire d'Océanographie de Villefranche, CNRS-Université Pierre et Marie Curie, B.P. 08, 06238 Villefranche-sur-mer, France

A R T I C L E I N F O A B S T R A C T

Article history: River floods and storm waves are major processes for the dispersal and deposition of terrestrial organic
Received 30 November 2007 matter (OM) in river-dominated coastal areas. A “wet storm” is connected to a flood with a high river
Received in revised form 5 April 2008 discharge, while a “dry storm” is not associated with a flood. To better understand the sedimentation
Accepted 17 April 2008
dynamics of terrestrial OM, especially soil OM, at the land–ocean interface during wet and dry storms, we
studied sediment trap and core material collected on the Têt inner shelf (NW Mediterranean) using multiple
Keywords:
organic proxies in combination with hydrodynamic parameters. The proportion of soil OM to the total OM in
NW Mediterranean
Soil organic matter
the trap material calculated based on the BIT (Branched and Isoprenoid Tetraether) index was higher during
Têt inner shelf both wet (~ 40%) and dry (~ 30%) storms than during non-storm periods (~ 10%). However, only surface
BIT index sediments (1-cm thick layers) recovered after December wet and moderate storms in 2003 showed enhanced
Wet storm soil OM percentages compared to the deeper sediments deposited during the last century. Given that the wet
Dry storm storm eroded 4-cm of seabed at the study site, flood-induced fresh soil OM was not deposited on the Têt
inner shelf during the wet storm. However, the following moderate storm caused resuspension of flood-
associated soil OM which was previously trapped nearshore. Accordingly soil OM was transported to the Têt
inner shelf and stored there until the dry storm occurred.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Therefore, the Têt system, a small river system characteristic of the
Mediterranean, behaves differently compared to larger systems, such
Continental margins are dynamic regions where most riverine as the Eel system in the northern Pacific coast, where ephemeral
particulate matter is deposited. Rivers associated with deltaic envi- (days–weeks) fine flood layers are deposited during wet storms (e.g.
ronments provide an important pathway to deliver terrestrial par- Mullenbach and Nittrouer, 2000; Wheatcroft and Borgeld, 2000; Fan
ticulate matter to the marine environment. Processes involved in et al., 2004).
transferring riverine particulate matter to the shelf are numerous: In order to study the effect of wet and dry storms on the OM
floods, storm waves, up and downwelling, cascading, eddies, internal deposition in coastal areas, we performed organic geochemical ana-
waves, etc (e.g. Dufau-Julliand et al., 2004; Fan et al., 2004; lyses on sediment trap samples collected during the storm events and
Palanques et al., 2006). River floods and storm waves are particularly surface sediments collected before and after each of these storm
important processes for the delivery, dispersal and deposition of events. We applied the newly developed BIT (Branched and
riverine particulate matter in river-dominated coastal oceans (e.g. Isoprenoid Tetraether) index as a proxy for soil organic matter (OM)
Ogston et al., 2000; Ahn and Grant, 2007). input. The branched GDGTs are derived from bacteria predominantly
A “wet storm” is connected to a flood with a high river discharge, thriving in soils (Weijers et al., 2006a,b) and transported via rivers to
while a “dry storm” occurs without significant river runoff (Ogston et the ocean margin (Hopmans et al., 2004; Herfort et al., 2006).
al., 2000). Sediment dynamics during wet and dry storms in the NW Previous studies carried out in the Gulf of Lions and on the Têt River
Mediterranean were studied (Guillén et al., 2006). An increase in (Kim et al., 2006, 2007) have shown that the branched GDGTs
suspended sediment concentration and erosion of seabed on the Têt detected in the Gulf of Lions were transported from land to sea by
inner shelf was noted during the wet storm of 4 December 2003. rivers. Therefore, branched GDGTs along with long-chain n-alkanes,
compounds derived from plant leaf wax, were used here as
⁎ Corresponding author. Tel.: +31 222 3695567; fax: +31 222 319674. continental lipid biomarkers. Conversely, crenarchaeol is mainly
E-mail address: jhkim@nioz.nl (J.-H. Kim). produced by non-extremophilic marine planktonic Crenarchaeota

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.04.019
J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238 229

Fig. 1. Map of study area showing the Gulf of Lions with main rivers and submarine canyons and detailed sedimentary map of the Têt prodelta (redrawn from the geological map of
Perpignan, 1/50,000e). The Têt POEM buoy site is located off Têt River mouth in 28 m water depth.

(Sinninghe Damsté et al., 2002), although this compound was also crenarchaeol was regarded as a marine lipid biomarker. Our approach
detected, albeit in small amounts, in terrestrial environments (Herfort based on multiple organic proxies in combination with hydrodynamic
et al., 2006; Kim et al., 2006, 2007; Weijers et al., 2006b). Therefore, parameters provided information on transport and depositional
230 J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

processes of riverine soil OM during wet and dry storms on the Têt lithology of the Têt watershed is dominated by old plutonic and
inner shelf. metamorphic rocks (granite, gneiss, mica schist and schist), which are
weathering resistant and only slowly mineralize (Salvayre, 1974; Garcia-
2. Study area Esteves et al., 2007). Limestone is present in the middle part of the
watershed. Further downstream, the river enters in its alluvial plain which
The Gulf of Lions is a river-dominated continental margin located in consists of thick fillings of the erosion products of the upstream rocks that
the NW Mediterranean between 42°N–3°E and 44°N–6°E (Fig. 1). Its accumulated during Pliocene and Quaternary periods. The adjacent shelf
crescent shaped continental shelf is fairly broad and its continental sediments are composed mostly of sands and silts between the shoreline
slope incised by numerous submarine canyons. The sediment and 20 m water depth and of clayey silts and silty clays between 20 m and
distribution displays a mid-shelf mud belt that follows the predomi- 40 m water depth (Aloïsi et al.,1973, Fig.1). The Têt River discharges water
nance of the Rhône River inputs and the general coastal circulation. The and sediment episodically onto the continental shelf, generating an
inner and outer shelf regions comprise mixed sandy mud deposits ephemeral prodeltaic deposit (e.g. Buscail et al., 1990, 1995).
(Aloïsi et al., 1973).
The Mediterranean rivers flowing into the Gulf of Lions (e.g. Têt, 3. Material and methods
Agly, Aude, Hérault) respond quickly to variations in rainfall with
episodic water discharges. In some areas, almost all annual rainfall falls 3.1. Meteorological data
within a few days (Ludwig et al., 2004). Therefore, the impact of storms
associated with a high precipitation on river discharge is largely The water discharge data of the Têt River in m3 s− 1 was obtained from
enhanced in the Mediterranean terrestrial basins. Consequently, the “HYDRO” data bank hosted at the French Ministry of Environment. The
transport of material from land to sea by rivers is strictly conditioned hourly water discharge data of Têt River represents the sum of the water
by synergic effects of storms and floods. discharge measured at the “Joffre” hydrologic station and at the Basse River,
The Têt system, a typical Mediterranean river system, is located in the a small affluent. Since suspended particulate matter (SPM) records for the
south-western part of the Gulf of Lions (NW Mediterranean, Fig. 1). The considered time periods in this study were not directly measured, a
Têt River catchment area is about 1400 km2 (Ludwig et al., 2004). The previously established relationship (Bourrin et al., 2006) between SPM in

Fig. 2. HPLC/APCI-MS base peak chromatogram of GDGT lipids in the trap sample (I-6) and GDGT structures: I, II, and III indicate the branched GDGTs and IV is crenarchaeol.
J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238 231

mg L− 1 and instantaneous river discharge in m3 s− 1 (Qi) was applied to the Sediment cores were recovered by divers using transparent Perspex
hourly water discharge data for the study periods to obtain SPMs in mg L− 1: tubes (7 cm diameter and 20 cm long) in close vicinity of the Têt buoy
site. Cores were obtained on 26 November 2003, 12 December 2003, 11
logSPM ¼ 0:294llogQi2 −0:0951logQi þ 0:9839ðr2 ¼ 0:60; n ¼ 1382Þ ð1Þ February 2004, and 15 March 2004. The cores were sliced at 1-cm
intervals. The trap and core samples were freeze-dried and homo-
The SPM discharges of the Têt River were further estimated based genized prior to analysis.
on the calculated SPMs (mg L− 1) and the measured river discharge data
(m3 s− 1) in ton per hour (T h− 1). 3.4. Geochemical analyses

3.2. Hydrodynamic data Bulk sediment and lipid analyses were done as previously
described by Kim et al. (2006). In brief, freeze-dried, acidified
Temperature and salinity were measured at 1 m above bottom samples were analyzed for total organic carbon concentrations (TOC)
(mab) with an Aanderaa RCM9 installed on a tripod frame at the Têt with an automatic CN-analyzer LECO 2000 at CEFREM. The precision
buoy site. The Têt buoy site is located 3 km off the Têt River mouth at for TOC was 0.2 wt.%. For stable carbon isotopic compositions of TOC
28 m water depth and is part of the POEM-L2R (Observation Platform (δ13Corg), acidified samples with 2 M HCl to remove carbonates were
of the Mediterranean Environment of the Littoral of Languedoc- combusted using a Flash EA 1112 Elemental Analyser interfaced with
Roussillon) station maintained by CEFREM since 2003. a ThermoFinnigan DeltaPlus mass spectrometer at NIOZ. Isotope
values were calibrated to a benzoic acid standard (δ13Corg = − 27.8‰
3.3. Sampling with respect to Vienna Pee Dee Belemnite (VPDB) standard calib-
rated on NBS-22 and corrected for blank contribution. The δ13Corg
A sequential cylindrical sediment trap (136 cm high with a filter dia- values are reported in the standard delta notation relative to VPDB
meter of 15.4 cm) mounted on a tripod frame was deployed at the Têt buoy standard. The analyses were done at least in duplicate. The analytical
site (28 m water depth, 42.7041°N, 3.0668°E) from 27 November 2003 to 16 reproducibility was ± 0.2‰.
December 2003 (12 bottles, 2 day collecting interval) and from 5 February Total lipid contents of sediments were measured at CEFREM using
2004 to 10 March 2004 (12 bottles, 3 day collecting interval). During the a colorimetric method (Barnes and Blackstock, 1973) after extraction
first deployment, only cups 4 to 6 were recovered in good condition while with a mixture of chloroform:methanol (CHCl3:MeOH, 2:1 v/v).
others were either lost during recovery, or heavily contaminated by Absorption of the products resulting from the colorimetric reaction
sediment deposited on the trap carousel. During the second deployment, between extract acidified with H2SO4 and a solution of Vanillin/H3PO4
the whole series of cups 1–12 were recovered in good condition. was measured at 520 nm with a Beckman spectrophotometer. For lipid

Fig. 3. Meteorological, hydrodynamic, and sedimentological parameters: hourly water discharge of the Têt River (m3 s− 1), hourly Têt River SPM discharge (T h− 1), and near-bottom
wave orbital velocity (m s− 1) from Guillén et al. (2006), temperature (°C), salinity, and SPM concentration (g L− 1) at 1 mab at the Têt buoy site.
232 J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

analysis, freeze-dried samples were extracted with an Accelerated performance liquid chromatography/atmospheric pressure positive
Solvent Extractor (DIONEX ASE 200) using a mixture of dichloro- ion chemical ionization mass spectrometry (HPLC/APCI-MS). GDGTs
methane (DCM):MeOH (9:1 v/v). The internal standard (Huguet et al., were detected by single ion monitoring of their [M + H]+ ions (dwell
2006) for GDGT quantifications was added into the total extracts time 237 ms) and quantification of the GDGT compounds was
before they were separated into three fractions over an Al2O3 column achieved by integrating the peak areas and using a C46 GDGT internal
using hexane:DCM (9:1 v/v), hexane:DCM (1:1 v/v), and DCM:MeOH standard according to Huguet et al. (2006). Values of BIT index were
(1:1 v/v), respectively. For aliphatic hydrocarbon (n-alkane) analysis, calculated according to Hopmans et al. (2004):
the hexane:DCM (9:1 v/v) fractions of the total lipid extract were
further purified over an AgNO3-impregnated silicagel column using BITindex ¼ ½I þ II þ III=½I þ II þ III þ IV ð2Þ
hexane as the eluent. The internal standard (perdeutero-n-C24 alkane)
was added into the purified fractions before GC analysis and used as
reference for n-alkane quantification. Analyses were performed on a The roman numerals refer to the GDGTs indicated in Fig. 2. I, II, and III
Hewlett Packard 5890 series II gas chromatograph. For GDGT analysis, are branched GDGTs and IV the isoprenoid GDGT, crenarchaeol. The
the DCM:MeOH (1:1 v/v) fractions were analyzed with a high reproducibility in the determination of the BIT index was better than 0.01.

Fig. 4. Results of bulk sediment and lipid analyses from Têt buoy site trap sediments: total mass flux (g m− 2 d− 1), TOC flux (g m− 2 d− 1), δ13Corg (‰ VPDB), ΣALK27–31 flux (mg m− 2 d− 1),
branched GDGT flux (μg m− 2 d− 1), crenarchaeol flux (μg m− 2 d− 1), and BIT index. TOC content (wt.%), total lipid (mg g−sed 1
), ΣALK27–31 (μg g−sed
1
), sum of branched GDGTs concentration
(ng g−sed
1
), and crenarchaeol concentration (ng g−sed
1
) are indicated as red coloured lines. Filled triangles indicate dates when sediment cores were taken in the Têt buoy site.
J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238 233

Fig. 5. Results of geochemical analyses from Têt buoy sediment cores: TOC contents (wt.%), δ13Corg (‰ VPDB), total lipids (mg g−sed
1
), ΣALK27–31 (μg g−sed
1
), BIT index, sum of branched
GDGTs concentration (ng g−sed
1
), and crenarchaeol concentration (ng g−sed
1
).
234 J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

4. Results following it. SPM concentration at the Têt buoy site drastically
increased during the wet and dry storms and to a lesser degree during
4.1. Meteorological and hydrodynamic parameters the moderate storms.

Têt River water and SPM discharge, near-bottom wave orbital 4.2. Sedimentology and bulk geochemistry
velocity, temperature, salinity, and suspended particulate matter
concentration data are shown in Fig. 3. For the considered time Total mass flux, TOC content and flux, and δ13Corg values of the
periods when sediment traps were deployed on the Têt buoy site, the particulate matter collected with the sediment trap deployed at the
hourly Têt River water discharges ranged from 5 to 363 m3 s− 1, with study site are shown in Fig. 4. During the major and moderate storm
two major peaks on 4 December 2003 at 12:00 am and on 21 February events that occurred in December 2003, the total mass fluxes were
2004 at 3:00 am, respectively. The peak water discharges were 363 2050 g m− 2 d− 1 and 1030 g m− 2 d− 1, respectively. During the dry storm
and 78 m3 s− 1, respectively. Floods in the Têt River are defined as a event in February 2004, the total mass flux drastically increased up to
river discharge N100 m3 s− 1 (Garcia-Esteves, 2005). Therefore, only a value of 1710 g m− 2 d− 1. During the non-storm period, the sediment
the first peak in water discharge on 4 December 2003 can be regarded fluxes were relatively low, varying from 2 to 340 g m− 2 d− 1.
as a flood. It lasted for two days and supplied 2⁎104 tons of sediment to The TOC content of the particulate matter was highest during the
the Têt prodelta area. December moderate storm. The TOC fluxes during the wet and dry storms
Based on the near-bottom current speed produced by surface were much higher (ca. 50 and 30 g m− 2 d− 1, respectively) than during
waves (near-bottom wave orbital velocity), two major (N1.2 m s− 1) and non-storm periods, varying between 0.1 g m− 2 d− 1 and 4 g m− 2 d− 1.
two moderate (0.9 m s− 1) storm events were observed during the δ13Corg values of the particulate matter during the wet and moderate
study period (cf. Guillén et al., 2006). The major storm event occurred storms (ca. −27‰) were significantly lower than those (ca. −25‰) during
on 4 December 2003 and was accompanied by a flood, resulting in a the dry storm.
“wet storm” event. However, the second major storm occurring on 21 The bulk parameter data (TOC and δ13Corg) of the four sediment cores
February 2004 was not related to a flood, and therefore it is described are plotted in Fig. 5. In general, TOC values were below 1 wt.%. However,
as a “dry storm” event although Têt River water and sediment the top 1 cm layers of the cores taken in December 2003 and February
discharges were slightly increased. Two moderate storms took place 2004 showed increased values, with 3.2 wt.% and 1.7 wt.%, respectively.
on 8 December 2003 and 14 March 2004. During these moderate δ13Corg values from the upper 5 cm varied substantially between −26.3‰
storms, increased Têt River water and particulate matter discharges and −22.8‰, while they were fairly constant below 5 cm.
were not observed.
Temperature measured at the Têt buoy site abruptly dropped 4.3. Lipid geochemistry
(about 1 °C) during the wet storm, while there was no major change
during the following moderate storm. In general, temperatures were The lipid data acquired from the sediment trap deployed at the Têt
lower (~ 4 °C) in February and March 2004 than in November and buoy site are shown in Fig. 4. The concentrations of total lipid (mg g−sed
1
),
December 2003. During the dry storm, temperature also dropped but ΣALK27–33 (μg g−sed
1
), and branched GDGTs (ng g−sed
1
) were highest during
less than during the wet storm (~0.5 °C). A significant salinity drop the December moderate storm and were lower during the following
was recorded after the December wet and the moderate storm February dry storm. In contrast, crenarchaeol concentration (μg g−sed 1
)

Fig. 6. Calculated terrestrial OM percentages based on δ13Corg and BIT index using a two end-member model and marine, soil, and plant OM percentages using a three end-member
model for Têt trap samples. Filled triangles indicate dates when sediment cores were taken in the Têt buoy site.
J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238 235

Fig. 7. Calculated terrestrial OM percentages based on δ13Corg and BIT index using a two end-member model in the upper panel and marine, soil, and plant OM percentages using a
three end-member model in the lower panel for Têt sediment core samples.

was much higher during non-storm periods than during storm events. 2003 and February 2004. The branched GDGTs and crenarchaeol
Due to limited amounts of trap material, only four samples could be concentrations basically followed the same pattern as the total lipid and
measured for the total lipid concentration and their flux values varied ΣALK27–33 concentrations. The BIT values were relatively constant (0.3 ±
between 0.1 and 0.9 g m− 2 d− 1, showing higher fluxes during storm events, 0.02) throughout the cores, except for the higher BIT values (up to 0.55)
although only one sample from a non-storm period was measured. The in the top 1 cm layers of the cores taken in December 2003 and
flux of the summed C27 to C31 n-alkanes (ΣALK27–33 flux) varied between February 2004.
0.02 and 11 mg m− 2 d− 1, with higher fluxes during storm events. The flux
of branched GDGTs, derived from soils, showed the same pattern as the 5. Discussion
ΣALK27–33 flux, varying between 1 and 135 μg m− 2 d− 1. In general, the
crenarchaeol fluxes were higher than those of the branched GDGTs. The 5.1. Source and abundance of terrestrial OM in Têt sediment traps
crenarchaeol fluxes ranged from 6 and 286 μg m− 2 d− 1. The BIT values of
the trap samples varied between 0.08 and 0.36, with generally higher In a previous study by Kim et al. (2007), Têt River water was
values during storm events than during intermediate non-storm periods. sampled at the station Perpignan from November 2000 to April 2002
The lipid data acquired from four sediment cores are shown in Fig. 5. and at the station Villelongue from January 2005 to February 2006.
The total lipid concentrations ranged from 0.1 to 1.7 mg g−sed 1
, with the The mean stable carbon isotopic composition of particulate organic
highest concentration in the sediment (0.5–1.0 cm) taken in December carbon (δ13CPOC) of the Têt River flowing to the Gulf of Lions was on
2003. ΣALK27–33 concentrations were fairly constant throughout four average −26.1‰ which is typical of C3 plant vegetation (e.g. Tyson,
cores, except for the top 1 cm layers of the cores taken in December 1995; Meyers, 1997) and depleted in 13C compared to that of seawater
236 J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

sampled in the Têt prodelta (−22.7‰), the latter being, close to typical OMmar, OMterr, and OMplant are the relative amount of marine, soil,
marine δ13CPOC values (−18 to −22‰, e.g. Meyers, 1994, 1997 and and plant OM, respectively. The end-member values of δ13C and BIT are
references cited therein). The δ13Corg values of the trap samples as follows: δ13Cmar =−20‰ (Fry and Sherr, 1984), δ13Csoil = −26‰ (Powers
studied here (−26.6 to −24.9‰) suggest the presence of substantial and Schlesinger, 2002), δ13Cplant =−28‰ (Hedges et al.,1986), BITsoil =0.85
amounts of terrestrial OM in the Têt traps. The δ13Corg values found (Kim et al., 2007), BITmar =0.00 (Kim et al., 2007), BITplant =0.00 (Hopmans
during the early December wet storm (4 December 2003) and the et al., 2004; Weijers et al., 2006b). These calculations show that the
following moderate storm (8 December 2003) were lower (− 26.6‰ estimated percentages of terrestrial OM (comprising both soil and plant
and −26.5‰, respectively) than that from the February dry storm OM) are quite substantial relative to those of marine OM (67 to 92%,
(−25.2‰), suggesting enhanced terrestrial OM contributions during Fig. 6). Furthermore, a substantial part of total OM is comprised of soil
the December storms. The BIT values of the trap samples (0.08 to 0.36) OM during the storm events (31 to 42%, Fig. 6).
were intermediate between those of the Têt River (0.85, Kim et al.,
2007) and of seawater (0.00, Kim et al., 2007), indicating a substantial 5.2. Source and distribution of terrestrial OM in Têt inner shelf sediment
input of soil OM into the sediment trap, especially during the storms.
Based on the δ13Corg values, we estimated the relative proportions The surface sediments (1-cm thick layers) recovered after the wet
of marine and terrestrial OM with a two end-member model, and moderate storms in December 2003 showed enhanced TOC, BIT
assuming terrestrial (δ13Cterr = −27‰) and marine (δ13Cmar = −20‰) values and lipid concentrations and lower δ13Corg values compared to
OM end-members (cf. Kim et al., 2006): those from the surface sediments collected before the wet storm and
after the dry and moderate storms. This indicates that the terrestrial
TOCsample ¼ OMmar þ OMterr ð3Þ OM fraction was enhanced in the surface sediments sampled in
December 2003 and February 2004. Terrestrial OM estimates based on
δ13Corg and BIT values using a two end-member model (see above)
δ13 Csample ⁎TOC ¼ δ13 Cmar ⁎OMmar þ δ13 Cterr ⁎OMterr ð4Þ
confirmed increases in terrestrial OM proportions after the wet and
moderate storms in December 2003 compared to those before the
where OMmar and OMterr are the contents of marine and terrestrial December wet storm in 2003 as well as after the February dry storm
OM, and δ13Cmar and δ13Cterr are the isotopic compositions of marine and the March moderate storm in 2004 (Fig. 7). The estimated
and terrestrial OM, respectively. The estimated contribution of percentages of soil OM using a three end-member model (see above)
terrestrial OM varied between 89 and 94% in December 2003 and suggest that a large part of total OM is comprised of soil OM (up to
between 70 and 74% in February and March 2004 (Fig. 6). During the 65%) after the wet and moderate storms than before the wet storm
December wet and moderate storms, terrestrial OM contribution was and after the dry and moderate storms (28 to 33%; Fig. 7).
higher than that during the February dry storm. Differences in bulk and lipid parameters were also observed below
Similar to the calculation of terrestrial OM based on the δ13Corg the sediment surface layers (Fig. 5). For example, sediments from 1 to
values, we estimated the portions of soil and marine OM, assuming soil 4 cm depth were depleted in TOC (0.2 wt.%) and total lipids (0.05–
(BITsoil = 0.85) and marine (BITmar = 0.00) OM end-members (Kim et al., 0.07 mg g−sed 1
) after the February dry and March moderate storms,
2007). The estimated contributions of soil OM based on the BIT indices compared to the lower sediments down to 10 cm (TOC 0.5–0.6 wt.%.
ranged from 39 to 42% in December 2003 and between 9 and 31% in and total lipids 0.12–0.14 mg g−sed1
). Likewise, considerable variations
February and March 2004 (Fig. 6). The BIT approach suggested relatively were recorded in δ Corg. In general, the δ13Corg values of the layers
13

higher amounts of soil OM during the December wet and moderate from 1 to 4 cm were enriched in 13C with an average value of −23.8‰,
storms than during the February dry storm. The trend in soil OM while the lower parts down to 10 cm had the average value of −25.0‰.
abundance is similar to that of terrestrial OM based on δ13Corg values. This suggests relatively higher abundance of terrestrial OM in the
The estimated relative terrestrial OM amounts based on δ13Corg lower parts of the sediment cores possibly because terrestrial OM is
values were substantially higher than the soil OM estimates based on better preserved (e.g. Hoefs, et al., 2002; Burdige, 2007). The layers of
BIT values (see Fig. 6). Likely, the δ13Corg comprises both soil and higher δ13Corg values were accompanied by slightly lower TOC and
vegetation OM. In addition, ancient refractory sedimentary OM, which total lipids contents but no significant changes were observed in other
has a depleted 13C value (e.g. Blair et al., 2003; Hwang et al., 2005), may lipid parameters.
be a significant component of suspended loads in small mountainous Estimated terrestrial OM percentages based on the δ13Corg values
rivers and some marine sediments. However, there are no sedimentary varied between 40% and 80%. This is at least twice as high as soil OM
rocks in the Têt watershed containing high percentages of organic percentages based on the BIT values, ranging from 20 to 40% (Fig. 7). In
carbon which could have contributed to the high terrestrial OM present day hydrodynamic conditions, flood-associated fine sedi-
portions observed in the Têt trap samples. A sewage treatment plant is ments freshly deposited do not remain in place more than 2 months
located above the Têt river station (see Fig. 1). Therefore, anthropo- on the Têt prodelta area (Bourrin et al., 2007). Therefore, δ13Corg
genic OM input possibly also influenced the relative terrestrial OM approaches seem to overestimate the proportion of terrestrial OM on
estimations based on the δ13Corg. However, it is impossible at this stage the Têt inner shelf. As mentioned above, this may lie partly in the
to estimate the portion of this source due to the lack of data. assumed end-member values of δ13Corg. However, the differences
We thus combine the BIT index with the δ13Corg to estimate relative between estimated terrestrial OM based on the δ13Corg and soil OM
proportions of marine, soil, and plant OM using a three end-member based on the BIT index are likely largely due to the large OM
model in a similar way as conducted by Gordon and Goñi (2003): contribution from contemporary terrestrial vegetations such as lignin
(e.g. Hedges and Mann, 1979; Da Cunha et al., 2001).
TOCsample ¼ OMmar þ OMsoil þ OMplant ð5Þ The trends of estimated contributions of plant OM using a three
end-member model were significantly different compared to those of
soil OM. In general, the layers between 1 cm and 4 cm contained lower
δ13 Csample ⁎TOCsample ¼ δ13 Cmar ⁎OMmar þ δ13 Csoil ⁎OMsoil plant OM percentages compared to the lower part down to 10 cm,
þ δ13 Cplant ⁎OMplant ð6Þ while soil OM percentages were fairly constant. This implies that
relatively labile, 13C enriched organic compounds of terrestrial OM
were selectively remineralized in the lower part (4–10 cm) of the
BITsample ⁎TOCsample ¼ BITmar ⁎OMmar þ BITsoil ⁎OMsoil sediment cores on the Têt inner shelf. Studies by Wang et al. (1998)
þ BITplant ⁎OMplant ð7Þ and Hwang et al. (2005) showed that in general lipids are more
J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238 237

refractory than other organic compound classes, e.g. total hydro- decrease of sediment supply by rivers due to climate change and/or
lysable amino acid (THAA) and total carbohydrate (TCHO). In their human activities (irrigation, damming, e.g. the Vinca damm on the Têt
studies, THAA and TCHO have higher δ13C values than those of lipids. River was built in 1975) may have induced a winnowing of fine
This suggests that selective degradation of THAA and TCHO could have material and thus a coarsening of upper sediment layers on the Têt
caused lower δ13C values in the lower part of sediment cores on the inner shelf.
Têt inner shelf, which in turn increased the estimated portions of plant
OM percentages. 6. Conclusions

5.3. Deposition of terrestrial OM on the Têt inner shelf Our study performed on the Têt inner shelf supports that the BIT
index is a useful proxy to trace flood-associated storm events in river-
The wet storm of early December was associated with the highest dominated continental margins such as the Gulf of Lions in the NW
water and sediment discharges of the Têt River into the prodelta and Mediterranean and to better understand riverine soil OM transport
the highest suspended sediment concentration and total mass flux in and depositional processes from rivers to oceans. The Têt inner shelf
the sediment trap within the time frame of this study (Fig. 3). did not accumulate terrestrial OM directly during the wet storm, when
However, Guillén et al. (2006) and Bourrin et al. (2007) observed that erosion was dominant at the study site but during the next moderate
during this December wet storm there was no flood deposit but storm, the flood-associated terrestrial OM was resuspended nearshore
erosion on the Têt inner shelf. The fresh water intrusion on the Têt and transported to the Têt inner shelf where it was stored until the dry
inner shelf related to this wet storm can be traced by an abrupt drop of storm occurred. As a result, the small Têt system acts as a bypass area
salinity which occurred after the wet storm. This suggests that fresh of terrestrial OM supplied by rivers mainly during the wet storm-
riverine OM was not delivered to the Têt buoy site during the wet associated floods and redistributed during following storms. Conse-
storm but only after the storm. The higher flux of terrestrial OM during quently, our study supports the previous results by Guillén et al.
the wet storm was due to wave resuspension of terrestrial OM which (2006) that the small Têt system behaves differently compared to
was deposited nearshore and on the Têt inner shelf prior to the larger systems where during wet storms a pluricentimetric flood layer
December wet storm rather than direct flood flux occurred during the deposition is frequently observed on the inner shelf. The BIT index
December wet storm. may be more useful to trace past flood events in a larger system where
Four hours after the peak of the wet storm, a 1-cm thick fluffy flood the sediment deposition is continuous due to high sedimentation
layer was accumulated on the Têt inner shelf and the moderate storm that rates in contrast to the small Têt system.
occurred after the wet storm caused about 4-cm thick deposition of
resuspended flood-associated sediment (Guillén et al., 2006). This in- Acknowledgements
dicates that enhanced terrestrial OM signals of the top 1-cm layer of the
sediment core sampled in December 2003 and February 2004 (Fig. 7) were We would like to thank G. Jeanty, J. Carbonne, and G. Saragoni at
not direct flood deposit but rather flood-associated storm deposit. The CEFREM, E. Hopmans, J. Ossebaar, and M. Kienhuis at NIOZ, and J. Guillén
Rhône River has high mean SPM discharge of about 7–10⁎106 106 T yr− 1 at CSIC for the sample preparation and analytical support. We ac-
(Sempéré et al., 2000; Pont et al., 2002), accounting for ~80% of the riverine knowledge constructive reviews by two anonymous reviewers. This
inputs to the Gulf of Lions (Durrieu de Madron et al., 2000). Considering study was supported by a Marie Curie Intra-European Fellowship grant
this large contribution of terrestrial OM through the Rhône River, the high to J.-H. Kim and the EUROSTRATAFORM project funded by the European
terrestrial OM layer on the Têt inner shelf was probably formed both by the Union.
supply from the Têt River and from other rivers located northwards of the
Têt River, e.g. Aude, Hérault, Rhône Rivers (Fig. 1). Especially, the floods References
associated to one single SE windstorm event (wet storm) are generally
concomitant in the whole Gulf of Lions. For example during the December Ahn, J.H., Grant, S.B., 2007. Characteristics of storm runoff and sediment dispersal in the
San Pedro Channel, southern California. Water Sci. Technol. 55, 519–526.
wet storm in 2003, all the rivers flowing into the Gulf of Lions generated a Aloïsi, J.-C., Got, H., Monaco, A., 1973. Garte géologique du précontinent languedocien au
big flood and especially the Rhône River delivered large amounts of water 1/250000ième. International Institute for Aerial Survey and Earth Sciences, the
(10,000 m3 s− 1, Palanques et al., 2006) and terrestrial material to the Netherlands.
Antonelli, C., Eyrolle, F., Rolland, B., Provansal, M., Sabatier, F., 2008. Suspended
coastal zone (Antonelli et al., 2008). Such terrestrial material was trapped
sediment and 137Cs fluxes during the exceptional December 2003 flood in the
nearshore and resuspended and further transported southwards during Rhône River, southeast France. Geomorphology 95, 350–360.
the moderate storm, forming high terrestrial OM layer on the Têt inner Barnes, H., Blackstock, J., 1973. Estimation of lipids in marine animals and tissue:
detailed investigations of the sulfovanilin method for total lipids. J. Exp. Mar. Biol.
shelf (Guillén et al., 2006).
Ecol. 12, 103–118.
When time passed, fine terrestrial OM on the Têt inner shelf was Blair, N.E., Leithold, E.L., Ford, S.T., Peeler, K.A., Holmes, J.C., Perkey, D.W., 2003. The
diluted by marine production and thus terrestrial OM proportion was persistence of memory: the fate of ancient sedimentary organic carbon in a modern
reduced on the Têt inner shelf. This dilution process can be traced in sedimentary system. Geochim. Cosmochim. Acta 67, 63–73.
Bourrin, F., Durrieu de Madron, X., Ludwig, W., 2006. Contribution of the study of coastal
the core top sediment taken in February 2004 which showed a lower rivers and associated prodeltas to sediment supply in North-western Mediterra-
proportion of terrestrial OM (88%) compared to that of the December nean Sea (Gulf of Lions). Life Environ. 56, 1–8.
core top sediment (95%). After the dry storm, no flood layer was Bourrin, F., Monaco, A., Aloïsi, J.-C., Sanchez-Cabeza, J.-A., Lofi, J., Heussner, S.,
Durrieu de Madron, X., Jeanty, G., Buscail, R., Saragoni, G., 2007. Last millennia
deposited and the moderate storm (15 March 2004) finally occurred sedimentary record on a micro-tidal low-accumulation prodelta (Têt NW Medi-
enhancing sediment erosion down to 1 cm depth. As a result, after the terranean). Mar. Geol. 243, 77–96.
dry and moderate storms, the core top from March 2004 has a rela- Burdige, D.J., 2007. Preservation of organic matter in marine sediments: controls,
mechanisms, and an imbalance in sediment organic carbon budgets? Chem. Rev.
tively lower proportion of terrestrial OM. 107, 467–485.
Besides the top 1-cm flood layers, the lower sediment layers con- Buscail, R., Pocklington, R., Daumas, R., Guidi, L., 1990. Fluxes and budget of organic
sisted of very-fine, well-sorted sands (Bourrin et al., 2007). Radio- matter in the benthic boundary layer over the northwestern Mediterranean
margin. Cont. Shelf Res. 10, 1089–1122.
carbon dating of T. communis shells of these layers and 210Pb dating
Buscail, R., Pocklington, R., Germain, C., 1995. Seasonal variability of the organic matter
showed similar ages from 1836 to 1950 (modern) and thus indicate in a sedimentary coastal environment: sources, degradation and accumulation
that these layers were formed not continuously during a period of (continental shelf of the Gulf of Lions-Northwestern Mediterranean Sea). Cont.
Shelf Res. 15, 843–869.
about one century. No layers with enhanced terrestrial OM were found
Da Cunha, L.C., Serve, L., Gadel, F., Blazi, J.-L., 2001. Lignin-derived phenolic compounds
related to catastrophic flooding of the last century (e.g. 1940) in the in the particulate organic matter of a French Mediterranean river: seasonal and
sedimentary records, suggesting this flood layer has been eroded. The spatial variations. Org. Geochem. 32, 305–320.
238 J.-H. Kim et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 228–238

Dufau-Julliand, C., Marsaleix, P., Petrenko, A., Dekeyser, I., 2004. Three-dimensionalmodel- Kim, J.-H., Ludwig, W., Schouten, S., Kerhervé, P., Herfort, L., Bonnin, J., Sinninghe Damsté, J.S.,
ing of the Gulf of Lion's hydrodynamics (northwest Mediterranean) during January 2007. Impact of flood events on the transport of terrestrial organic matter to the ocean: a
1999 (MOOGLI3 Experiment) and late winter 1999: Western Mediterranean Inter- study of the Têt River (SW France) using the BIT index. Org. Geochem. 38, 1593–1606.
mediate Water's (WIW's) formation and its cascading over the shelf break. J. Geophys. Ludwig, W., Serrat, P., Cesmat, L., Garcia-Esteves, J., 2004. Evaluating the impact of the recent
Res. 109. doi:10.1029/2003JC002019. stemperature increase on the hydrology of the Têt River (Southern France). J. Hydrol.
Durrieu de Madron, X., Abassi, A., Heussner, S., Monaco, A., Aloïsi, J.-C., Radakovitch, O., 289, 204–221.
Giresse, P., Buscail, R., Kerherve, P., 2000. Particulate matter and organic carbon Meyers, P.A., 1994. Preservation of source identification of sedimentary organic matter
budgets for the Gulf of Lions (NW Mediterranean). Oceanol. Acta 23, 717–730. during and after deposition. Chem. Geol. 144, 289–302.
Fan, S., Swift, D.J.P., Traykovski, P., Bentley, S., Borgeld, J.C., Reed, C.W., Niedoroda, A.W., Meyers, P.A., 1997. Organic geochemical proxies of paleoceanographic, paleolimnologic
2004. River flooding, storm resuspension, and event stratigraphy on the northern and paleoclimatic processes. Org. Geochem. 27, 213–250.
California shelf: observations compared with simulations. Mar. Geol. 210, 17–41. Mullenbach, B.L., Nittrouer, C.A., 2000. Rapid deposition of fluvial sediment in the Eel
Fry, B., Sherr, E.B., 1984. Delta-C-13 measurements as indicators of carbon flow in Canyon, northern California. Cont. Shelf Res. 20, 2191–2212.
marine and freshwater ecosystems. Contrib. Mar. Sci. 27, 13–47. Ogston, A.S., Cacchione, D.A., Sternberg, R.W., Kineke, G.C., 2000. Observations of storm
Garcia-Esteves, J., 2005. Transferts géochimiques en Méditerranée: exemple de la and river flood-driven sediment transport on the northern California continental
rivière Têt et de son bassin versant. PhD thesis, Université de Perpignan. shelf. Cont. Shelf Res. 20, 2141–2161.
Garcia-Esteves, J., Ludwig, W., Kerhervé, P., Probst, J.-L., Lespinas, F., 2007. Predicting the Palanques, A., Durrieu de Madron, X., Puig, P., Fabres, J., Guillén, J., Calafat, A., Canals, M.,
impact of land use on the major element and nutrient fluxes in coastal Mediterranean Heussner, S., Bonnin, J., 2006. Suspended sediment fluxes and transport processes
rivers: the case of the Têt River (Southern France). Appl. Geochem. 22, 230–248. in the Gulf of Lions submarine canyons: the role of storms and dense water
Gordon, E.S., Goñi, M.A., 2003. Sources and distribution of terrigenous organic matter cascading. Mar. Geol. 234, 43–61.
delivered by the Atchafalaya River to sediments in the northern Gulf of Mexico. Pont, D., Simonnet, J.P., Walter, A.V., 2002. Medium-term changes in suspended
Geochim. Cosmochim. Acta 67, 2359–2375. sediment delivery to the ocean: consequences of catchment heterogeneity and
Guillén, J., Bourrin, F., Palanques, A., Durrieu DeMadron, X., Puig, P., Buscail, R., 2006. river management (Rhône river, France). Estuar. Coast. Shelf Sci. 54, 1–18.
Sediment dynamics during wet and dry storm events on the Têt inner shelf (SW Powers, J.S., Schlesinger, W.H., 2002. Geographic and vertical patterns of stable carbon
Gulf of Lions). Mar. Geol. 234, 129–142. isotopes in tropical rain forest soils of Costa Rica. Geoderma 109, 141–160.
Hedges, J.I., Mann, D.C., 1979. The characterization of plant tissues by their lignin Salvayre, H., 1974. Care géologique des Pyrénées-Orientales (care géologique simplifiée
oxidation products. Geochim. Cosmochim. Acta 43, 1803–1807. à l'usage des élèves du ler et 2ème cycle et des amateurs de Géologie). Éditions
Hedges, J.I., Clark, W.A., Quay, P.D., Richey, J.E., Devol, A.H., Santos, U.D., 1986. Ophrys, Paris.
Compositions and fluxes of particulate organic material in the Amazon River. Sempéré, R., Charrière, B., Van Wambeke, F., Cauwet, G., 2000. Carbon inputs of the
Limnol. Oceanogr. 31, 717–738. Rhone River to the Mediterranean Sea: biogeochemical implications. Glob.
Herfort, L., Schouten, S., Boon, J.P., Woltering, M., Baas, M., Weijers, J.W.H., Sinninghe Biogeochem. Cycles 14, 669–681.
Damsté, J.S., 2006. Characterization of transport and deposition of terrestrial Sinninghe Damsté, J.S., Hopmans, E.C., Schouten, S., van Duin, A.C.T., Geenevasen, J.A.J.,
organic matter in the southern North Sea using the BIT index. Lim. Oceanogr. 51, 2002. Crenarchaeol: the characteristic core glycerol dibiphytanyl glycerol tetraether
2196–2205. membrane lipid of cosmopolitan pelagic crenarchaeota. J. Lipid Res. 43, 1641–1651.
Hoefs, M.J.L., Rijpstra, W.I.C., Sinninghe Damsté, J.S., 2002. The influence of oxic Tyson, R.V., 1995. Sedimentary Organic Matter: Organic Facies and Palynofacies.
degradation on the sedimentary biomarker record I: evidence from Madeira Chapman and Hall, New York.
Abyssal Plain turbidites. Geochim. Cosmochim. Acta 66, 2719–2735. Wang, X.-C., Druffel, E.R.M., Griffin, S., Lee, C., Kashgarian, M., 1998. Radiocarbon studies
Hopmans, E.C., Weijers, J.W.H., Schefuß, E., Herfort, L., Sinninghe Damsté, J.S., Schouten, S., of organic compound classes in plankton and sediment of the northeastern Pacific
2004. A novel proxy for terrestrial organic matter in sediments based on branched Ocean. Geochim. Cosmochim. Acta 62, 1365–1378.
and isoprenoidtetraether lipids. Earth Planet. Sci. Lett. 224, 107–116. Weijers, J.W.H., Schouten, S., Hopmans, E.C., Geenevasen, J.A.J., David, O.R.P., Coleman, J.M.,
Huguet, C., Hopmans, E.C., Febo-Ayala, W., Thompson, D.H., Sinninghe Damsté, J.S., Pancost, R.D., Sinninghe Damsté, J.S., 2006a. Membrane lipids of mesophilic anaerobic
Schouten, S., 2006. An improved method to determine the absolute abundance of bacteria thriving in peats have typical archaeal traits. Environ. Microbiol. 8, 648–657.
glycerol dibiphytanyl glycerol tetraether lipids. Org. Geochem. 37, 1036–1041. Weijers, J.W.H., Schouten, S., Spaargaren, O., Sinninghe Damsté, J.S., 2006b. Occurrence
Hwang, J., Druffel, E.R.M., Komada, T., 2005. Transport of organic carbon from the and distribution of tetraether membrane lipids in soils: implications for the use of
California coast to the slope region: a study of Δ14C and δ13C signatures of organic the TEX86 proxy and the BIT index. Org. Geochem. 37, 1680–1693.
compound classes. Glob. Biogeochem. Cycles 19 (GB2018). doi:10.1029/ Wheatcroft, R.A., Borgeld, J.C., 2000. Oceanic flood deposits on the northern California
2004GB002422. shelf: large-scale distribution and small-scale physical properties. Cont. Shelf Res.
Kim, J.-H., Schouten, S., Buscail, R., Ludwig, W., Bonnin, J., Sinninghe Damsté, J.S., 20, 2163–2190.
Bourrin, F., 2006. Origin and distribution of terrestrial organic matter in the NW
Mediterranean (Gulf of Lions): application of the newly developed BIT index.
Geochem. Geophys. Geosyst. 7 (Q11017). doi:10.1029/2006GC001306.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / p a l a e o

Geochemical evidence for high-resolution variations during deposition of the


Holocene S1 sapropel on the Cretan Ridge, Eastern Mediterranean
G. Gennari a,⁎, F. Tamburini b, D. Ariztegui c, I. Hajdas d, S. Spezzaferri a
a
University of Fribourg, Department of Geosciences, Earth Sciences, Ch. du Musée 6, 1700 Fribourg, Switzerland
b
Institute of Plant Nutrition, ETH Zurich, Eschikon 33, 8315 Lindau, Switzerland
c
University of Geneva, Section of Earth Sciences, Rue des Maraîchers 13, 1205 Geneva, Switzerland
d
Ion Beam Physics, PSI and ETH Zurich, Schafmattstrasse 20, 8093 Zurich, Switzerland

A R T I C L E I N F O A B S T R A C T

Article history: Major and minor element distributions and solid phase phosphorus contents were measured in a sediment
Received 25 January 2008 core from the Cretan Ridge that contains the Holocene S1 sapropel. Micro-XRF ultra-high resolution analysis
Received in revised form 6 June 2008 reveals multiple Mn peaks in the oxidized upper portion of the sapropel, which implies either a non-steady-
Accepted 10 June 2008
state upward remobilization of Mn or not constant downward diffusion of bottom water oxygen. Sequential
extraction allowed the identification of different phosphorus phases. Detrital phosphorus increases in the
Keywords:
Sapropel
sapropel layer, suggesting enhanced delivery from land during sapropel deposition. Higher organic carbon to
Chemical elements organic phosphorus ratios in the sapropel indicate enhanced regeneration of phosphorus relative to carbon
Phosphorus under low-oxygen conditions. Fourier analysis on the ultra-high resolution XRF data reveals frequencies that
Cycles are provisionally assigned to millennial to decennial solar cycles. Although these cycles have been reported
from several continental and marine archives, they have never been documented from Sapropel S1 in the
Eastern Mediterranean.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction 2) increased Nile river runoff linked to the periodic enhancement


of the African–Asian monsoons (Rossignol-Strick 1983, 1985);
The Eastern Mediterranean Sea is characterized by the occurrence 3) increased rainfalls and river discharge along the northern part of
of organic matter-rich layers, termed sapropels. According to the the Eastern Mediterranean Sea (Cramp et al., 1988; Rohling and
original definition by Kidd et al. (1978), sapropels are well-delimited Hilgen, 1991).
layers within open marine sediments, with thickness N1 cm and In the “increased productivity” model, sapropel deposition is
organic carbon content N2%. The relatively high organic carbon linked to enhanced organic matter flux (Calvert, 1983; Calvert et al.,
content of a sapropel is due to the combination of enhanced supply 1992), since the present production of organic matter in the eastern
and limited degradation of organic matter at the sediment–water Mediterranean cannot account for the high values of organic carbon
boundary, which is water-depth dependent (Murat and Got, 2000). (TOC) characterizing these layers (Calvert, 1983).
Thereby, the difference between the organic carbon content of a According to other authors (Castradori, 1993; Rohling, 1994; Emeis
sapropel and that of the surrounding sediments seems to be a more et al., 1998; Emeis et al., 2000), an increase of nutrient input via river
suitable parameter to define these layers (Murat, 1991; Ariztegui et al., runoff could have caused enhanced primary production and also
2000). promoted the stratification of the water column, coupling the two
Among the various processes that have been invoked to explain models. A significant increase of productivity at times of sapropel
sapropel formation (see Rohling, 1994 for a review), the “stagnation/ deposition is also revealed by paleo-productivity proxies, as Ba
anoxia” and the “increased productivity” models are the most and marine barite concentration (e.g., Thomson et al., 1995, 1999;
discussed. According to the stagnation/anoxia model, anoxic bottom Martinez-Ruiz et al., 2000, 2003; Gallego-Torres et al., 2007).
conditions are caused by a strong stratification of the water column Enhanced productivity has also been explained with a radical change
that prevents vertical mixing and oxygen supply to the bottom waters. in the Mediterranean Sea circulation, from the modern anti-estuarine
Different explanations have been proposed for the origin of this to an estuarine circulation (Sarmiento et al., 1988; Howell and Thunell,
stratification: 1) Eurasian ice-sheet melt water entering the Medi- 1992). Results of recent modeling studies on thermohaline circulation
terranean Sea (Olausson, 1961; Ryan, 1972; Aksu et al., 1995); suggest that a weakening of the present-day anti-estuarine circulation
can lead to the deposition of enough organic carbon to account, at
⁎ Corresponding author. Fax: +41 26 300 97 42. least, for the formation of the youngest sapropel S1 (Myers et al.,
E-mail address: giordana.gennari@unifr.ch (G. Gennari). 2000; Stratford et al., 2000). Uncertainties still exist regarding the

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.06.007
240 G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248

extension of the anoxic/dysoxic layer in the water column; this layer Table 1
has been described as a large water mass extending below the mixed Results of AMS 14C dating on planktonic foraminifers. Calibration using OxCal 3.10
(Bronk Ramsey, 2005)
layer (Murat and Got, 2000; Stratford et al., 2000), as well as an
“anoxic blanket” above the sediment/water interface (Casford et al., Lab code Depth Material 14
C age Calibrated age
2003). More recently, Bianchi et al. (2006) suggest that a modified (cm) (yr BP) (yr BP)
thermohaline circulation supplying oxygen only to the first 500 m of ETH-34228 1.0–2.0 Mixed planktonic foram 1050 ± 55 610 ± 70
the water column may be responsible for the development of an ETH-34229 22.0–22.5 Globigerinoides ruber 6165 ± 80 6570 ± 180
anoxic blanket at the sea-floor, when coupled with higher productiv- ETH-34230 44.5–45.5 Globigerinoides ruber 11910 ± 110 13405 ± 235
ity in the euphotic zone. ETH-34231 67.5–68.0 Globigerinoides ruber 17200 ± 160 19925 ± 375

Regardless of the mechanism that led to their formation, sapropel A reservoir effect of 400 years was subtracted.
occurrence is cyclic and is thought to be linked to minima in the
precession cycle, corresponding to Northern Hemisphere summer
insolation maxima, with a periodicity of about 21 kyr (Rossignol-
Strick, 1983, 1985; Hilgen, 1991; Lourens et al., 1996). 2.1. Organic matter
In this research, we present a high-resolution geochemical study of
sapropel S1, since any change in environmental or climatic conditions Organic matter characterization was performed at the University of
is reflected in major and minor element abundances (e.g., Wehausen Neuchâtel on about 80 mg of ground and sieved sediment with a Rock-
and Brumsack, 1998). Furthermore, because our proxy data offer a Eval 6, using the standard whole rock pyrolysis method (Espitalié
very detailed and continuous record, the identification of high-order et al., 1986).
cyclicities is attempted.
2.2. Phosphorus
2. Materials and methods
Phosphorus was extracted from the dried sediment after grinding
Gravity core SIN97-01GC was collected on the morphological high to b125 μm. A 5-step sequential extraction technique was used to
between Crete and the Peloponnesus (35°49′07″ N, 22°42′02″ E; distinguish among different phosphorus sedimentary phases (SEDEX
water depth 933 m) during the 15/97 cruise of R/V Urania (Fig. 1). method; Ruttenberg, 1992; Tamburini et al., 2002). Solid phases were
Sampling for this study was limited to the uppermost part of the core progressively dissolved by using a series of extractants, each chosen in
(section 1; 0.0–68.0 cm bsf), since it contains a thick sapropel S1 order to extract phosphorus associated with well defined sedimentary
overlaid by a few centimeters-thick oxidized interval, both inter- components: (1) loosely-bound P adsorbed on mineral surfaces,
bedded in a hemipelagic muddy succession. (2) iron and manganese oxy-hydroxides (referred in the text as CDB-
Sediment samples of 0.5 cm thickness for geochemical analyses P), (3) authigenic apatite, (4) detrital material (igneous, metamorphic
were taken along the core at 3 cm spacing; a contiguous sampling was and “aged” authigenic apatite), and (5) organic matter. The obtained
also performed at any change in sediment type or color. Sediments sample solutions were diluted in milli-Q water and a color developing
were then dried at room temperature and split into sub-samples. An agent was added to the solutions following the ascorbic acid method
aliquot of the material was gently ground in an agate mortar and for phosphate (Eaton et al., 1995). All the concentrations were then
passed through a sediment sieve (b125 μm), with the minimal extent measured using a Perkin Elmer UV/Vis Spectrophotometer Lambda 10
of grinding necessary to achieve this size. at the GEA lab (University of Neuchâtel).

Fig. 1. Location map of core SIN97-01GC.


G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248
Fig. 2. Corg (weight %) and elemental ratios (versus Al, Fe or Ca). Elemental ratios were smoothed using a 10 points mobile average in order to highlight the dominant trends in the data. Ox S1:oxidized sapropel S1; S1: visual dark-colored
sapropel S1.

241
242 G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248

Because of the reduction of Fe by dithionite and subsequent


complexation by citrate, during the second step of this extraction it is
possible to separate also ferric Fe (CDB-Fe). CDB-P and ferric Fe
concentration were measured using a ICP-OES at the Soil Chemistry
Institute of the ETH Zurich.

2.3. Elemental analyses

Ultra high-resolution qualitative analyses of major and minor


elements (Mn, Fe, Ca, Mg, Al, Sr, Cl, Ti) were performed on macroscopic
contiguous samples by X-ray microfluorescence (µ-XRF), using
an EDAX Eagle III XPL μprobe (40 kV, 400 μA, 50/55 Dtm) with an
analytical spot size of 50 μm. Average spacing between analytical
points was 0.35 mm. Two profiles were acquired on two different
sets of samples, giving consistent results. The µ-XRF elemental
analyses were carried at the Section of Earth Sciences of the University
of Geneva.

14
2.4. AMS C dating

Radiocarbon ages were obtained using accelerator mass spectro- Fig. 4. S2 versus Corg diagram for S1 samples. Boundaries between the three kerogen
metry (AMS) technique. The 14C dating was performed at the AMS type fields (dashed lines) from Langford and Blanc-Valleron (1990). Regression curve
facilities, ETH and PSI in Zurich (Table 1). Four selected samples (solid line) and migrated regression curve (dotted line) are also shown.
each containing 700 to 1000 specimens of planktonic foraminifers
(15–18 mg of pure calcium carbonate) were hand-picked under a
binocular microscope and then ultrasonically cleaned to remove 2.5. Fourier analysis
any kind of contamination and/or diagenetic alteration. Samples
were then dissolved in concentrated phosphoric acid (Hajdas et al.,
2004a). The resulting carbon dioxide was then converted into Four AMS 14C ages (calibrated) were used to infer sedimentation
graphite as described by Hajdas et al. (2004b). A reservoir age cor- rates and construct the chronological framework of the studied
rection of 400 years (Siani et al., 2000) was applied to the obtained sedimentary sequence. To identify potential high-frequency cycles
radiocarbon conventional ages, which were then calibrated using the within the sedimentary record, Fourier analysis was performed on
program OxCal 3.10 (Bronk Ramsey, 2005). the obtained µ-XRF elemental ratio dataset (average of 33 points
per cm). The Fourier analysis was carried out on selected elements
versus Al ratios by using the software Analyseries (Paillard et al.,
1996).

3. Results

3.1. Organic matter characterization by Rock-Eval

Organic carbon content in the studied core is generally low, but


shows a remarkable increase in correspondence with the visual dark-
colored sapropel (Fig. 2). Values range between 0.06 and 0.16 wt%
in the “normal” hemipelagic sediments and from 0.39 to 0.89 wt%
in sapropel S1.
Tmax (420 °C on average) and hydrogen index (HI) values suggest
that the organic material in sapropel S1 is mainly composed of
partially degraded marine organic matter. Low HI values (75–200 mg
HC/g C org; Fig. 3) could indicate a strong mineral matrix
effect (Espitalié et al., 1986). The S2 versus Corg diagram (Fig. 4)
confirms the presence of such matrix effect. In fact, according to
Langford and Blanc-Valleron (1990), the regression line in a S2
versus Corg diagram should pass through the axes origin, and a
positive x-intercept, as in this case, could indicate the presence of
rock-matrix effect. The high correlation degree of the samples in the
S2 versus Corg graph seems to indicate a common origin of the
organic material. In fact, if the matrix effect is eliminated by
migrating the regression line towards the axes origin, all the
sapropel S1 samples fall in the marine organic matter field (type II
kerogen field). Even though these results suggest that the main
component of the organic material characterizing sapropel S1 is of
marine origin, we do not exclude the contribution of terrestrial
material. The amount of TOC in samples below and above the visual
Fig. 3. Corg (wt%) concentration and HI (mg HC/g Corg), as determined by Rock-Eval, and sapropel (b0.16 wt%) is too low, so it is not possible to distinguish
Corg/Porg (molar ratio) curve. the effect of clay mineral absorption.
G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248 243

Fig. 5. Percentages of total P in the analyzed phosphorus sedimentary phases.

3.2. Phosphorus 3.4. 14C chronology

Phosphorus extracted from sediments is dominated by the Age assessment of core SIN97-01GC is based on four AMS 14C ages
authigenic phase (apatite) that accounts on average for the 45% of obtained from planktonic foraminifers. The shallower sample (1–2 cm
total phosphorus (calculated as the sum of the extracted phases; depth) was too small for monospecific dating, therefore mixed
Fig. 5). Both loosely-bound P and detrital P profiles (Fig. 6) show a planktonic foraminifers were picked. Both radiocarbon conventional
general down-core increasing trend, while the organic-bound P is ages (referred in the text as 14C yr) and calibrated ages (2σ error;
characterized by an opposite trend. Higher concentrations of detrital referred as cal yr BP) are shown in Table 1.
and, to a lesser extent, of organic-bound P are observed in the core The sediments immediately below the base of S1 sapropel
interval between 28 and 44.5 cm bsf, corresponding to the visual are dated at 11910 ± 110 14C yr BP (corresponding to 13405 ±
sapropel. The sapropel S1 is also characterized, in the uppermost part, 235 cal yr BP), while the sample immediately above the top of the
by a peak in CDB-P and in CDB-Fe. In this interval, CDB-P represents oxidized S1 displays an age of 6165 ± 80 14C yr BP (corresponding to
the most abundant solid phase (up to 37% of total P, Fig. 5). 6570 ± 180 cal yr BP). Although there is general agreement in con-
The molar Corg/Porg (Fig. 3) is higher in the sapropel (200–500) in sidering S1 to have been formed at around 9000–7000 years BP
comparison to the overlying and underlying hemipelagic sediments (see Fontugne et al., 1994 for a review), several authors reported
(25–130). The sediments below S1, however, show Corg/Porg values older ages for the onset of the conditions leading to the formation
slightly higher then the sediment above due to the general decreasing of this anoxic layer (e.g., Troelstra et al., 1991; Howell and Thunell,
trend of organic-bound P. 1992). In particular the age of the base of S1 in core SIN797-01GC
is consistent with the ones reported by Rohling et al. (1993) for core
3.3. Sediment geochemistry T87/2/20G (11680 14C yr BP; 13300 cal yr BP reported in Hayes et al.,
1999) and by Anastasakis (2007) for the south-eastern Mediterra-
The chemical characterization is presented as element/Al ratios nean Sea.
(Fig. 2) in order to compensate for carbonate dilution (Wehausen and The top of the visual sapropel has an inferred age of 7600 14C yr BP
Brumsack,1998). Fe, Mn and Al concentration curves are also shown for a (corresponding to 8340 cal yr BP), well in agreement with the widely
comparison (Fig. 7). The Mn/Al ratio shows a remarkable increase at the reported dating (Rohling, 1994; Mercone et al., 2000).
top of the visual sapropel S1, developing a series of well-distinguished
peaks in the oxidized part of the sapropel. The Fe/Al ratio shows a 3.5. Fourier analysis
positive shift in the uppermost part of S1 and in the above 5 cm. Cl/Al and
Ti/Al are both enriched in the core interval corresponding to the original The sedimentation rate was calculated by fitting a regression line
sapropel (S1 + Ox S1). Ca/Al and Mg/Al are covariant too, but with an of calibrated ages against depth (Fig. 8). The calculated sedimentation
opposite trend, and show depletion in the same interval. The Ca/Fe and rate is constant at 3.4 cm/kyr.
Ca/Ti ratios decrease in the whole original sapropel interval, while Sr/Ca The performed spectral analysis shows well-distinguished peri-
curve shows a positive shift in the same core interval. odicities with statistically significant peaks at 85–70, 60 and 50 years.
244
G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248
Fig. 6. Phosphorus solid phases content (in μmol P/g) and CDB-Fe (in μmol Fe/g). Ox S1:oxidized sapropel S1; S1: visual dark-colored sapropel S1.
G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248 245

increasing with increasing location depth, has been clearly established


by Murat and Got (2000). According with these authors, differences in
organic carbon concentrations are more likely the result of variations
in organic matter degradation than in organic matter availability, and
are linked to lower degradation rates or to higher burial efficiency at
deeper sites, resulting in increasing TOC content with depth. In this
framework, the low TOC content of S1 on the Cretan Ridge fits well
with the relatively shallow depositional environment (933 m).
As a consequence of the reduced preservation of organic matter,
also the HI values are relatively low, as shown by the Rock-Eval results;
this is in agreement with the positive linear correlation between HI
and TOC concentration as also reported by Bouloubassi et al. (1999) for
Pliocene–Pleistocene sapropels.

4.2. Phosphorus phases and C/P ratios

CDB-P and CDB-Fe show a peak at the top of S1, and not above the
visual sapropel, as it was expected (Slomp et al., 2002, 2004). This is
probably due to the presence of well-crystallized Fe oxides in the
oxidized sapropel (Slomp et al., 1996) that are a less efficient sink for
phosphorus in comparison with amorphous phases. The observed Fe
enrichment can then be attributed to dissolution of amorphous Fe
sulfides during the CDB step of the sequential extraction (Slomp et al.,
1996). Both CDB-P and CDB-Fe curves, however, resemble closely the
distribution shown by Slomp et al. (2002, 2004) suggesting that these
peaks could anyway be related to the redox front.
Although the detrital P fraction extracted by the SEDEX method is
operationally defined (Ruttenberg, 1992), and the extraction solution
used (1 M HCl) might also dissolve sedimentary phases (e.g., clay and
Fe minerals) that have resisted previous steps, we are confident that
this fraction well represents detrital input. In fact, the amount of P
contained in clays is negligible (Ruttenberg, 1992), and dithionite and
citrate (step II) are the most efficient extractants for Fe oxides (Slomp
et al., 1996). Generally, only well well-crystallized oxides might persist
after this step, but these would probably also be of detrital origin.
Previous studies have shown that the detrital P well correlates
with other detrital proxies, such as quartz and ice rafted debris
(e.g. Tamburini et al., 2002, 2003) which were independently mea-
sured. Considering that, and knowing that detrital P is not involved in

Fig. 7. Fe, Mn and Al concentrations (counts/second). All curves were smoothed using a
10 points mobile average in order to highlight the dominant trends in the data. Ox S1:
oxidized sapropel S1; S1: visual dark-colored sapropel S1.

Millennial and centennial periodicities are also observed though not


significant above the 95% confidence limit (see Fig. 9).

4. Discussion

4.1. TOC content and organic matter preservation

The visual dark dark-colored S1 sapropel from core SIN97-01GC is


characterized by a TOC content distinctively higher than in the
uppermost and lowermost sediments. The displayed values, ranging
from 0.39 to 0.89 wt%, are however well below the 2% organic carbon
content originally used to define these anoxic layers (Kidd et al., 1978).
Indeed, lower TOC contents have been widely reported for sapropel S1
in the Eastern Mediterranean (e.g., Murat and Got, 2000). A linear
relationship between TOC percentages and water depth, with values Fig. 8. Calibrated 14
C ages versus depth. Linear regression line is plotted.
246 G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248

Fig. 9. Fourier analysis on Mg/Al, Mn/Al, Ca/Al, Sr/Al, Ti/Al and Fe/Al curves (B-Tuckey Spectrum, Bartlett window. Bandwidth: 0.0002522. Error estimation on Power spectrum:
0.546215 b ΔP / P b 2.55813). Sedimentation rates obtained by the regression curve of the calibrated ages versus depth plot.

biological P cycling, we consider its variations mainly as an indication as corroborated by our results and widely reported for the Eastern
of changes in detrital supply. In this case, its distribution indicates an Mediterranean (e.g., Emeis et al., 1996), we can use a C/P Redfield ratio
increase of detrital input during sapropel S1 deposition, as also of 106–117 (Redfield et al., 1963; Anderson and Sarmiento, 1994) as a
suggested by the slight increase of Ti/Al ratio in the core interval reference value for the original C/P ratio. In sapropel S1, Corg/Porg
corresponding to the original sapropel. values are higher than the expected Redfield ratio, ranging from 200
Organic-bound P does not positively correlate with organic C. to 500, while they display lower values in the overlying and un-
Because the organic matter in the sapropel is mainly of marine origin, derlying sediments (b 130). Our data are consistent with an enhanced
G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248 247

regeneration of P in respect to C during sapropel deposition (Ingall Considering the known influence of changes in insolation in the
and Jahnke, 1994; Slomp et al., 2002). The difference between the Mediterranean and northern African region and their direct link to
observed Corg/Porg values and the one reported by Slomp et al. (2002; sapropel formation (e.g. Rossignol-Strick, 1985), the presence of these
Corg/Porg = 300–1085 mol/mol) is consistent with the poor organic cycles confirms the strong sensitivity of Mediterranean climate even
matter preservation evident in the low TOC and HI values. at a decennial scale.

4.3. Post-depositional alteration of sapropel S1 5. Conclusions

The geochemical signature of sapropel S1 from the Cretan Ridge is Our high-resolution geochemical study of sapropel S1 from the
consistent with an active post-depositional oxidation due to the Cretan Ridge (Eastern Mediterranean) shows that:
transition between different bottom water oxygenation levels at the
end of the sapropel deposition (e.g., Thomson et al., 1999). Mn/Al 1) The top of the original sapropel was affected by post-depositional
peaks, and to a lesser extend the enrichment of Fe/Al observed in the alteration by a downward penetrating oxidation front, resulting in
oxidized sapropel, are linked to the downward penetration of the the removal of organic matter. The ultra high-resolution µ-XRF
oxidation front (Van Santvoort et al., 1996), that led to post- elemental analysis reveals a more articulated Mn/Al profile than
depositional removal of organic matter, as evidenced in the upper- expected, with the development of a multiple peak pattern. This
most part of the original sapropel. Nevertheless the Mn/Al profile peculiar profile may be explained by a “not constant” upward
shows a more articulated shape than the characteristic double peaks migration of dissolved Fe and Mn and/or by a non-constant
so far widely described (overview in Thomson et al., 1999). Van downward diffusion of bottom water oxygen, probably linked to
Santvoort et al. (1996) already illustrated the importance of sampling rapid variations in the sedimentation rate.
resolution in order to detect early diagenetic signals. As a case study, 2) The Corg/Porg ratio is in agreement with an enhanced regeneration
they showed the different shape of the solid phase Mn profile at low (1 of phosphorus in respect to C caused by low low-oxygen content of
peak; from 1 to N2.5 cm resolution) and high high-resolution (0.5 cm) bottom waters and sediments (Ingall and Jahnke, 1994) during
sampling (2 peaks) (boxcore BC 12, Eastern Mediterranean). With a sapropel deposition (Slomp et al., 2002).
ultra-high resolution (0.3 mm), our Mn/Al curve shows multiple peaks 3) Though aware of a possible analytical bias, we consider detrital
that may derive from a non-constant upward flux of dissolved Fe2+ phosphorus as a reliable indicator of detrital supply from the
and Mn2+ and/or from a non-constant downward diffusion of bottom continent. In this case, its increase supports the hypothesis of an
water oxygen, probably linked to rapid and short-lasting variations in enhanced detrital input during sapropel S1 deposition.
sedimentation rates. 4) The very high-resolution proxy record obtained by XRF analysis
The Fe/Al curve presents a broad positive shift with respect to allowed us to perform a spectral analysis. High-frequency
background values, because of increased amount of Fe-oxy-hydro- millennial to decennial-scale solar cycles centered at 1700, 550,
xides (Ox S1) and of pyrite or other amorphous Fe-sulfates (from the 210, 85–70, 60 and 50 years were identified, suggesting that also
top of the visual sapropel down to about 35 cm). during sapropel S1 deposition, climate in the Mediterranean region
The Cl/Al profile may provide further evidence of the sapropel was paced by solar variability even at short periodicities. Our
original thickness, being linked to the higher porosity that usually approach suggests that the use of µ-XRF scanning technique is a
characterizes organic-rich sediments (Thomson et al., 1995). powerful tool that can be used in order to obtain high high-
The negative shift in Ca content with respect to both Fe and Sr can resolution records suitable for the identification of meaningful
be related to differences in calcite preservation and abundance of Sr- periodicities.
rich aragonite, in S1 and in normal pelagic sediments.
Acknowledgements
4.4. High-frequency cyclicity
We warmly thank Cesare Corselli and Daniela Basso for providing
Millennial to decennial-scale periodicities have been observed at the core material, Thierry Adatte for the Rock-Eval analyses and Kurt
very different locations, in both marine and continental archives (e.g., Barmettler for ICP-OES measurements. We also thank Phil Meyers and
Hulu cave, China; GISP2 and GRIP ice cores, Greenland; monsoon two anonymous reviewers for their useful revision of the manuscript.
climate in Africa) and using different proxy records (see Tyson et al., This research is funded by the Swiss National Foundation (Ref.
2002). Their origin is certainly related to solar activity (Bond et al., 200021-111694).
2001), but the mechanisms of the complex interplay between sun and
climate are still under debate (Beer et al., 2000; Bond et al., 2001;
Turney et al., 2005). References
Also in core SIN97-01GC, the Fourier analysis performed on the
Aksu, A.E., Yasar, D., Mudie, P.J., 1995. Paleoclimatic and paleoceanographic conditions
ultra-high resolution elemental ratio records reveals the presence of leading to development of sapropel layer S1 in the Aegean Sea. Palaeogeogr.
well-distinguished millennial to decennial cyclicities centered at Palaeoclimatol. Palaeoecol. 116, 71–101.
around 1700, 550, 210, 85–70, 60 and 50 years (Fig. 9). The cycles Anastasakis, G., 2007. Clay mineral distribution patterns in the southeastern
Mediterranean Sea during the late Quaternary. Geol. Carpath. 58 (4), 383–395.
detected in our records at 210 yr (Suess Cycles) and 85–70 yr Anderson, L.A., Sarmiento, J.L., 1994. Redfield ratios of remineralization determined by
(Gleissberg Cycles) are most likely related to solar variability (see nutrients data analysis. Glob. Biogeochem. Cycles 8 (1), 65–80.
Chambers et al., 1999 for a review). The 550 years periodicity has been Ariztegui, D., Asioli, A., Lowe, J.J., Trincardi, F., Vigliotti, L., Tamburini, F., Chondrogianni, C.,
Accorsi, C.A., Bandini Mazzanti, M., Mercuri, A.M., van der Kaars, S., McKenzie, J.A.,
detected in the Holocene GISP2 ice core (Greenland; Stuiver et al., Oldfield, F., 2000. Palaeoclimatic reconstructions and formation of sapropel S1:
1995) and in North Atlantic sediment records (Chapman and inferences from Late Quaternary lacustrine and marine sequences in the Central
Shackleton, 2000), giving indication of an interconnection between Mediterranean region. Paleogeogr. Paleoclimatol. Palaeoecol. 158, 215–240.
Beer, J., Mende, W., Stellmacher, R., 2000. The role of the sun in climate forcing. Quat. Sci.
atmospheric and oceanic variability (Chapman and Shackleton, 2000).
Rev. 19, 403–415.
According to other authors (e.g., Debret et al., 2007) solar forcing could Bianchi, D., Zavatarelli, M., Pinardi, N., Capozzi, R., Capotondi, L., Corselli, C., Masina, S.,
be a more probable explanation for these periodicities. 2006. Simulations of ecosystem response during the sapropel S1 deposition event.
The 1700 years periodicity corresponds to the ∼1800 yr frequen- Paleogeogr. Paleoclimatol. Palaeoecol. 235, 265–287.
Bond, G., Kromer, B., Beer, J., Muscheler, R., Evans, M.N., Showers, W., Hoffman, S., Lotti-
cies already identified in lake sediment records, and it has been Bond, R., Hajdas, I., Bonani, G., 2001. Persistent solar influence on North Atlantic
attributed to variation in insolation (Skilbeck et al., 2005). climate during the Holocene. Science 294, 2130–2136.
248 G. Gennari et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 239–248

Bouloubassi, I., Rullkötter, J., Meyers, P.A., 1999. Origin and transformation of organic Mercone, D., Thomson, J., Croudace, I.W., Siani, G., Paterne, M., Troelstra, S., 2000.
matter in Pliocene–Pleistocene Mediterranean sapropels: organic geochemical Duration of S1, the most recent sapropel in the eastern Mediterranean Sea, as
evidence reviewed. Mar. Geol. 153, 177–197. indicated by accelerator mass spectrometry radiocarbon and geochemical
Bronk Ramsey, C., 2005. Improving the resolution of radiocarbon dating by statistical evidence. Paleoceanography 15 (3), 336–347.
analysis. In: Levy, T.E., Higham, T.F.G. (Eds.), The Bible and Radiocarbon Dating: Murat, A., 1991. Enregistrement sédimentaire des paléonvironments quaternaries en
Archaeology, Text and Science. Equinox, London, pp. 57–64. Méditerranée Orientale. Thése de doctorat, Université de Perpignan, 280 pp.
Calvert, S.E., 1983. Geochemistry of the Pleistocene sapropels and associated sediments Murat, A., Got, H., 2000. Organic carbon variations of the eastern Mediterranean
from the Eastern Mediterranean. Oceanol. Acta 6, 225–267. Holocene sapropel: a key for understanding formation processes. Palaeogeog.
Calvert, S.E., Nielsen, B., Fontugne, M.R., 1992. Evidence from nitrogen isotope ratios for Palaeoclimatol. Palaeoecol. 158, 241–257.
enhanced productivity during formation of eastern Mediterranean sapropels. Myers, P.G., Haines, K., Rohling, E.J., 2000. Modelling the paleocirculation of the
Nature 359, 223–225. Mediterranean: the last glacial maximum and the Holocene with emphasis on the
Casford, J.S.L., Rohling, E.J., Abu-Zied, R.H., Fontanier, C., Jorissen, F.J., Leng, M.J., formation of sapropel S1. Paleoceanography 13 (6), 586–606.
Schmiedel, G., Thomson, J., 2003. A dynamic concept for eastern Mediterranean Olausson, E., 1961. Studies of deep sea cores. Reports of the Swedish Deep-Sea
circulation and oxygenation during sapropel formation. Palaeogeogr. Palaeoclima- Expedition, 1947–1948, Vol.8, pp. 323–438.
tol. Palaeoecol. 190, 103–119. Paillard, D., Labeyrie, L., Yiou, P., 1996. Macintosh program performs time-series
Castradori, D., 1993. Calcareous nannofossils and the origin of eastern Mediterranean analysis. Eos Trans. AGU 77 (379).
sapropels. Paleoceanography 8, 459–471. Redfield, A.C., Ketchum, B.H., Richards, F.A., 1963. The influence of organisms in the
Chambers, F.M., Ogle, M.I., Blackford, J.J., 1999. Palaeoenvironmental evidence for solar composition of sea water. In: Hill, M.N. (Ed.), The Sea. Interscience, New York, pp. 26–77.
forcing of Holocene climate: linkages to solar science. Prog. Phys. Geogr. 23, 181–204. Rohling, E.J., 1994. Review and new aspects concerning the formation of eastern
Chapman, M.R., Shackleton, N.J., 2000. Evidence of 550-year and 1000-year cyclicities in Mediterranean sapropels. Mar. Geol. 122, 1–28.
North Atlantic circulation patterns during the Holocene. Holocene 10 (3), 287–291. Rohling, E.J., Hilgen, F.J., 1991. The eastern Mediterranean climate at time of sapropel
Cramp, A., Collins, M., West, R., 1988. Late Pleistocene–Holocene sedimentation in the formation: a review. Geol. Mijnb. 70, 253–264.
NW Aegean Sea: a paleoclimatic–paleoceanographic reconstruction. Palaeogeogr. Rohling, E.J., Jorissen, F.J., Vergnaud-Grazzini, C., Zachariasse, W.J., 1993. Northern
Palaeoclimatol. Palaeoecol. 68, 61–77. Levantine and Adriatic Quaternary planktic foraminifera; reconstruction of
Debret, M., Bout-Roumazeilles, V., Grousset, F., Desmet, M., McManus, J.F., Massei, N., paleoenvironmental gradients. Mar. Micropaleontol. 21, 191–218.
Sebag, D., Petit, J.R., Copard, Y., Trenteseaux, A., 2007. The origin of the 1500-year Rossignol-Strick, M., 1983. African monsoon, an immediate response to orbital
climate cycles in Holocene North-Atlantic records. Clim. Past 3, 569–575. insolation. Nature 304, 46–49.
Eaton, A.D., Clesceri, L.S., Greenberg, A.E., 1995. Standard Methods for the Examination of Rossignol-Strick, M., 1985. Mediterranean Quaternary sapropels, an intermediate
Water and Wastewater, 19th Ed. American Public Health Association, Washington, D.C. response of the African monsoon to variation of insolation. Palaeogeog. Palaeocli-
Emeis, K.C., Robertson, A.H.F., Richter, C., et al., 1996. Proc. ODP. Init. Repts., 160. Ocean matol. Palaeoecol. 49, 237–263.
Drilling Program, College Station, TX. Ryan, W.B.F., 1972. Stratigraphy of Late Quaternary sediments in the Eastern
Emeis, K.C., Schulz, H.M., Struck, U., Sakamoto, T., Doose, H., Erlenkeuser, H., Howell, M., Mediterranean. In: Stanley, D.J. (Ed.), The Mediterranean Sea: A natural sedimenta-
Kroon, D., Paterne, M., 1998. Stable isotopes and temperature records of sapropels tion laboratory. Dowden, Hutchinson & Ross, Stoudsburg, PA, pp. 149–170.
from ODP Sites 964 and 967: constraining the physical environment of sapropel Ruttenberg, K.C., 1992. Development of a sequential extraction method for different
formation in the Eastern Mediterranean Sea. In: Robertson, A.H.F., Emeis, K.-C., forms of phosphorus in marine sediments. Limnol. Oceanogr. 37 (7), 1460–1482.
Richter, C., Camerlenghi, A. (Eds.), Proc. ODP, Sci. Results, 160. Ocean Drilling Sarmiento, J.L., Herbert, T., Toggweiler, J.R., 1988. Mediterranean nutrient balance and
Program, College Station, TX, pp. 309–331. episodes of anoxia. Glob. Biogeochem. Cycles 2, 427–444.
Emeis, K.C., Struck, U., Schulz, H.M., Rosenberg, R., Bernasconi, S., Erlenkeuser, H., Siani, G., Paterne, M., Arnold, M., Bard, E., Métivier, B., Tisnerat, N., Bassinot, F., 2000.
Sakamoto, T., Martinez-Ruiz, F., 2000. Temperature and salinity variations of Radiocarbon reservoir ages in the Mediterranean Sea and Black Sea. Radiocarbon 42 (2),
Mediterranean Sea surface waters over the last 16.000 years from records of 271–280.
planktonic stable oxygen isotopes and alkenone unsaturation ratios. Palaeogeogr. Skilbeck, C.G., Rolph, T.C., Hill, N., Woods, J., Wilkens, R.H., 2005. Holocene millennial/
Palaeoclimatol. Palaeoecol. 158, 259–280. centennial-scale multiproxy cyclicity in temperate eastern Australian estuary
Espitalié, J., Deroo, G., Marquis, F., 1986. La pyolyse Rock-Eval et ses applications — III sediments. J. Quat. Res. 20 (4), 327–347.
partie. Rev. Inst. Fr. Pet. 41 (1), 73–89. Slomp, C.P., Van der Gaast, S.J., Van Raaphorst, W., 1996. Phosphorus binding by poorly
Fontugne, M.R., Amold, M., Labeyrie, L., Paterne, M., Calvert, S.E., Duplessy, J.C., 1994. crystalline iron oxides in North Sea sediments. Mar. Chem. 52, 55–73.
Paleoenvironment, sapropel, chronology and Nile river discharge during the last Slomp, C.P., Thomson, J., De Lange, G., 2002. Enhanced regeneration of phosphorous
20,000 years as indicated by deep-sea sediment records in the eastern during formation of the most recent eastern Mediterranean sapropel (S1). Geochim.
Mediterranean. In: Bar-Yosef, O., Kra, R.S. (Eds.), Late Quaternary Chronology and Cosmochim. Acta 66 (7), 1171–1184.
Paleoclimate of the eastern Mediterranean. Radiocarbon, pp. 75–88. Slomp, C.P., Thomson, J., De Lange, G., 2004. Controls on phosphorus regeneration and
Gallego-Torres, D., Martinez-Ruiz, F., Paytan, A., Jiménez-Espejo, F.J., Ortega-Huertas, M., burial during formation of eastern Mediterranean sapropels. Mar. Geol. 203, 141–159.
2007. Pliocene–Holocene evolution of depositional conditions in the eastern Stratford, K., Williams, R.G., Myers, P.G., 2000. Impact of the circulation on sapropel
Mediterranean: role of anoxia vs. productivity at time of sapropel deposition. formation in the eastern Mediterranean. Glob. Biogeochem. Cycles 14 (2), 683–695.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 246, 424–439. Stuiver, M., Grootes, P.M., Braziunas, T.F., 1995. The GISP2 δ18O record of the past
Hajdas, I., Bonani, G., Zimmerman, S.H., Mendelson, M., Hemming, S., 2004a. C-14 ages 16,500 years and the role of the sun, ocean and volcanoes. Quat. Res. 44, 341–354.
of ostracodes from Pleistocene lake sediments of the western Great Basin, USA— Tamburini, F., Huon, S., Steinmann, P., Grousset, F.E., Adatte, T., Föllmi, K.B., 2002.
Results of progressive acid leaching. Radiocarbon 46, 189–200. Dysaerobic conditions during Heinrich event 4 and 5. Evidence from phosphorus
Hajdas, I., Bonani, G., Thut, J., Leone, G., Pfenninger, R., Maden, C., 2004b. A report on distribution in a North Atlantic deep-sea core. Geochim. Cosmochim. Acta 66 (23),
sample preparation at the ETH/PSI AMS facility in Zurich. Nucl. Inst. Methods Sec. B 4069–4083.
223–224, 267–271. Tamburini, F., Föllmi, K.B., Adatte, T., Bernasconi, S.M., Steinmann, P., 2003. Sedimentary
Hayes, A., Rohling, E.J., De Rijk, S., Kroon, D., Zachariasse, W.J., 1999. Mediterranean phosphorus record from the Oman margin: new evidence of high productivity
planktonic foraminiferal faunas during the last glacial cycle. Mar. Geol. 153, 239–252. during glacial periods. Paleoceanography 18 (1), 1015.
Hilgen, F.J., 1991. Extension of the astronomically calibrated (polarity) time scale to the Thomson, J., Higgs, N.C., Wilson, T.R.S., Croudace, I.W., De Lange, G.J., van Santvoort, P.J.M.,
Miocene/Pliocene boundary. Earth Planet Sci. Lett. 107, 349–368. 1995. Redistribution and geochemical behaviour of redox-sensitive elements around
Howell, M.W., Thunell, R.C., 1992. Organic carbon accumulation in Bannock Basin: S1, the most recent eastern Mediterranean sapropel. Geochim. Cosmochim. Acta 59
evaluating the role of the productivity in the formation of eastern Mediterranean (17), 3487–3501.
sapropels. Mar. Geol. 103, 461–471. Thomson, J., Mercone, D., De Lange, G., van Santvoort, P.J.M., 1999. Review of recent
Ingall, E., Jahnke, R., 1994. Evidence for enhanced phosphorus regeneration from marine advances in the interpretation of eastern Mediterranean sapropel S1 from geochem-
sediments overlain by oxygen depleted waters. Geochim. Cosmochim. Acta 58 (11), ical evidences. Mar. Geol. 153, 77–89.
271–275. Troelstra, G., Gansenn, G.M., Van Der Borg, K., De Jong, A.F.M., 1991. A late Quaternary
Kidd, R.B., Cita, M.B., Ryan, W.B.F., 1978. Stratigraphy of eastern Mediterranean sapropel stratigraphic framework for eastern Mediterranean sapropel S1 based on AMS 14C
sequences recovered during DSDP Leg 42A and their paleoenvironmental dates and stable oxygen isotopes. Radiocarbon 33 (1), 15–21.
significance. In: Hsü, K.J., Montadert, L., et al. (Eds.), init. Rep. DSDP, 423 (pt. 1). Turney, C., Baillie, M., Clemens, S., Brown, D., Palmer, J., Pilcher, J., Reimer, P., Leuschner, H.H.,
US Government Printing Office, Washington, DC, pp. 421–443. 2005. Testing solar forcing of pervasive Holocene cycles. J. Quat. Sci. 20 (6), 511–518.
Langford, F.F., Blanc-Valleron, M.M.,1990. Interpreting Rock-Eval pyrolysis data using graphs Tyson, P.D., Cooper, G.R.J., McCarthy, T.S., 2002. Millenial to multi-decadal variability of
of pyrolizable hydrocarbons vs. total organic carbon. A.A.P.G. Bul. 74 (6), 799–804. the climate of Southern Africa. Int. J. Climatol. 22, 1105–1117.
Lourens, L.J., Antonarakou, A., Hilgen, F.J., van Hoof, A.A.M., Vergnaud-Grazzini, C., Van Santvoort, P.J.M., De Lange, G.J., Thomson, J., Cussen, H., Wilson, T.R.S., Krom, M.D.,
Zachariasse, W.J., 1996. Evaluation of the Plio-Pleistocene astronomical time-scale. Ströhle, K., 1996. Active post-depositional oxidation of the most recent sapropel (S1)
Paleoceanography 11, 391–413. in sediments of the Eastern Mediterranean Sea. Geochim. Cosmochim. Acta 60 (21),
Martinez-Ruiz, F., Kastner, M., Paytan, A., Ortega-Huertas, M., Bernasconi, S.M., 2000. 4007–4024.
Geochemical evidence for enhanced productivity during S1 sapropel deposition in Wehausen, R., Brumsack, H.J., 1998. The formation of eastern Mediterranean sapropels:
the eastern Mediterranean. Paleoceanography 15, 200–209. constraints from high-resolution major and minor element studies. In: Robertson,
Martinez-Ruiz, F., Paytan, A., Kastner, M., Gonzalez-Donoso, J.M., Linares, D., Bernasconi, S.M., A.H.F., Emeis, K.-C., Richter, C., Camerlenghi, A. (Eds.), Proc. ODP, Sci. Results, 160.
Jimenez-Espejo, F.J., 2003. A comparative study of the geochemical and mineralogical Ocean Drilling Program, College Station, TX, pp. 207–217.
characteristic of the sapropel S1 in the western and eastern Mediterranean. Palaeogeog.
Palaeoclimatol. Palaeoecol. 190, 23–37.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Role of sea-level forced sedimentary processes on the distribution of organic


carbon-rich marine sediments: A review of the Late Quaternary sapropels in the
Mediterranean Sea
Rossella Capozzi a,⁎, Alessandra Negri b
a
Dipartimento di Scienze della Terra e Geologico-Ambientali, University of Bologna, Via Zamboni 67, I-40126 Bologna, Italy
b
Dipartimento di Scienze del Mare, Polytechnic University of Marche, Via Brecce Bianche, I-60131 Ancona, Italy

A R T I C L E I N F O A B S T R A C T

Article history: The 124 ka to present sapropel S5–S1 series in the eastern Mediterranean Sea, correspond to the last Glacial–
Received 3 December 2007 Interglacial period, and permit detailed investigation of 5th order sedimentary frequencies. The Sapropel
Received in revised form 5 May 2008 layers are clustered in high-frequency Milankovian cycle that occur within sequence stratigraphic 4th order
Accepted 6 May 2008
cycles of late Pleistocene-Holocene sea-level oscillations. The sea-level changes had an amplitude exceeding
120 m and have been correlated to the glacial-interglacial cycles modulated by the 100 ka eccentricity. Using
Keywords:
Sapropel
published data we examined occurrences of the complete S5–S1 series in the Mediterranean Sea and then we
Sea level focussed on two stratigraphic sequences in the central Adriatic shelf and Middle Adriatic Depression where
Sequence stratigraphy sapropels S5 and S1 were deposited during periods of sea-level rise recorded by the Transgressive Systems
Productivity Tract (Marine Isotopic Stage 5 and 1, respectively). Sapropels S4 and S3 were deposited during the Highstand
Mediterranean Sea Systems Tract, that developed throughout MIS 5 of the late Pleistocene sequence. Sapropel S2 should have
been deposited during the warm MIS 3.3 and followed the MIS 4 sea-level drop when sediment supply was
enhanced during the late highstand and falling stage of sea-level of the late Pleistocene sequence. The
pronounced MIS 2 cold stage led to a lowstand and to the regional erosional surface (Sequence Boundary).
We also compared the Adriatic sapropel S5–S2 sequence (4th order) to the Mid-Pliocene cluster “O” sequence
(3rd order). In both cases, sapropels are well formed before the increase of sediment supply due to
progradation of the shelf wedge, that occurs in the late high stand and initial falling stage in sea-level. Finally,
we considered the influence of sea-level lowering which could cause a weakening of North Adriatic Deep
Water formation, therefore influencing the Mediterranean deep water circulation. This condition would
favour oxygen depletion, but it does not correspond to a well developed sapropel (as in the case of S2). Based
on the assumption that productivity increases related to insolation play a dominant role in their
sedimentation, sapropel S2 is the result of low amplitude insolation maxima, that were unable to generate
a productivity increase leading to the burial of conspicuous amount of organic matter. Therefore, in the
Adriatic Sea the absence of the S2 sapropel is correlated to the high sedimentation rate diluting the signal,
whereas in the eastern Mediterranean basin this sapropel is virtually absent because of a weak orbital forcing
combined with oxidation that erased its record.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction explain their mode of sedimentation. Initial ideas invoked stagnation


as the principal cause (e.g., Olausson, 1961; Cita et al., 1977; Vergnaud-
Sapropels are dark, organic matter rich, often laminated sediments Grazzini et al., 1977; Thunell and Williams, 1989; Aksu et al., 1995).
that have cyclically been deposited in the Mediterranean area starting However, other authors considered enhanced primary production in
from the late Miocene. The study of these sediments has been the the photic zone to be primary among the factors controlling the
subject of debate during the last 50 years and the literature regarding concentration and accumulation of organic matter in the sea floor, and
the argument is very abundant (see Negri et al., this volume). hence the major cause of sapropel deposition (Calvert, 1983; Pedersen
Sapropels were defined by Kidd et al. (1978) as well-delimited and Calvert, 1990 among others).
layers within open marine sediments, with thickness N1 cm and More recently, a general agreement has emerged that sapropel
organic carbon content N2%. Two models have been forwarded to deposition is accompanied by conditions of anoxia or dysoxia (e.g.
Emeis et al., 1996; Cramp and O’Sullivan, 1999) and that the recurrence
⁎ Corresponding author. Fax: +39 0512094522. of these layers is related to the Milankovitch precessional cycles
E-mail address: rossella.capozzi@unibo.it (R. Capozzi). (Hilgen, 1991), favouring a combination of productivity and stagnation

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.05.009
250 R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257

(e.g. Rohling and Gieskes, 1989; Rohling, 1994; Emeis et al., 2000a,b; Relation between sea-level oscillations and the Corg-rich levels
Warning and Brumsack, 2000). Although both factors have evolved from have been described mainly in third order sequences (1–3 My
the Pliocene to Pleistocene (Passier et al.,1999; Martinez-Ruiz et al., 2003) according to Vail et al., 1977; Posamentier et al., 1988). Within these
sapropel formation in the eastern Mediterranean is evidently related to sequences each Systems Tract, which is constituted by genetically
variations in the intensity of the African monsoon. During Earth axis associated stratigraphic units that were deposited during specific
precession minima, corresponding to maximum summer insolation in phases of the relative sea-level cycle (Posamentier et al., 1988), can
the northern hemisphere, the increased monsoonal precipitation over include higher frequency cycles (fourth order: 0.1–1 My; 5th order:
northern Africa enhanced runoff to the Mediterranean Sea (e.g., 0.01–0.1 My; Mitchum and Van Wagoner, 1991) that correspond to the
Rossignol- Strick, 1983, 1985, 1999; Rossignol-Strick et al., 1982). The Milankovian orbital perturbations. The climatic oscillations that
increased fresh water input lowered the salinity of the surface waters generate the sapropels are related to eccentricity, obliquity and
causing the slow down or arrest of thermohaline ventilation in the deeper precession and therefore corresponds to cycles of 4th and 5th order
parts of the Mediterranean and leading to anoxic bottom water. At the (Mitchum and Van Wagoner, 1991). Therefore, for a detailed study of
same time, the increased runoff combined with a shallowing of the the relation between the sedimentary record and deposition of the
nutricline into the photic zone led to enhanced primary productivity and Corg-rich levels we must focus attention on those sedimentary
therefore an increased export of organic material to the sea floor (Rohling successions in which these cycles are well developed. This situation
and Gieskes, 1989; Rohling, 1991; Corselli et al., 2002). happens especially in the late Pleistocene, where large amplitude
The extensive literature regarding sapropels is mainly based on glacial-interglacial oscillations of sea-level generate complete 4th
studies regarding micropaleontological and geochemical aspects, order sequences (Vail et al., 1977; Mitchum and Van Wagoner, 1991).
whereas studies of the sedimentological aspect are generally rare
(cf. Sigl et al. 1978; Stow et al. 2001; Hassold et al., 2003). Furthermore, 2. Sapropel record in the Mediterranean sea
very little is known about the possible effect of sea-level variations on
sapropel deposition. The sapropels that formed during times of low We focussed on a short stratigraphic interval between Marine
sea-level stand are especially interesting as the low sea-level at these Isotopic Stages (MIS) 5 and 1, which corresponds to the last 124 ka. In
times likely restricted circulation and exchange through the Gibraltar the Mediterranean Sea this interval records two large sea-level
and Sicily straits (Zahn et al., 1987) and also drastically reduced the oscillations (over 120 m) and a continuous sedimentary record.
size of the Adriatic Sea, known today as a source of major formation of The synchronous occurrence of sapropels within different deposi-
Eastern Mediterranean deep water. tional systems and from various physiographic settings is useful to
Literature dealing with the role of sea-level variations on understand the most favourable conditions for their record in the
deposition of Cenozoic sapropels is not abundant and so far limited sedimentary succession. With this purpose, we made an inventory of all
to the Middle Pliocene (Capozzi and Picotti, 2003; Roveri and Taviani, the sites in the Mediterranean basin, where a continuous sedimentary
2003; Marsaglia et al., 2004; Capozzi et al., 2006), whereas it is more succession of the last 124 ka occurs and the complete series from
abundant for older successions, generally due to the fact that Paleozoic sapropel S1 to sapropel S5 is available.
and Mesozoic black shales are objects of hydrocarbon exploration. Table 1 summarizes for each location (ODP sites, DSDP sites and
A few papers have documented that some Neogene sapropels, other published data, see Fig. 1) the presence of the “Brunhes”
Cretaceous OAEs and older black shales occurred during trangressions sapropels (Cita et al., 1977) and/or the occurrence of other horizons
and high sea-level (Wignall, 1994; Stefani 2002; Capozzi and Picotti not clearly labelled.
2003; Arthur and Sageman 2005; Beckmann et al. 2005), whereas In the Western Mediterranean and in the Tyrrhenian Sea the
others sequences have been correlated to low stands in sea-level “Brunhes” sapropels do not occur (Table 1) although, Murat (1999)
(Cobianchi and Picotti, 2001; Henz and Zeng, 2003). recognized some minor organic carbon-rich horizons in the Alboran
According to the conceptual model by Huc (1988), organic matter Sea, their occurrences are in the last interglacial and do not
accumulation results from the combination of several favoring factors correspond to the S1 sapropels found in the Eastern Mediterranean.
that are, first of all, high primary productivity (Pedersen and Calvert In the Eastern Mediterranean Basin (Fig. 1, Table 1), occurrences of
1990) and enhanced preservational conditions. The accumulation is the “Brunhes” sapropel (Cita et al., 1977) are widespread, although the
modulated by other factors intrinsic to sedimentation rate, such as series are sometimes locally incomplete. In particular, concentrating
dilution, time of residence of the organics at the sediment surface, on the last two interglacials, the most recognizable horizons are S5
water depth, detrital or biogenic mineral phases and distance from the and S1, whereas sapropel S2 was recognized only in cores RC9-181,
shore (Stow et al., 2001). ODP 967, M40/1-22 (Vergnaud-Grazzini et al., 1977; Emeis et al., 1996;
From this conceptual model, the aim of our paper is to assess the Löwemark et al., 2006); in addition, S3 and S4 do not appear in some
role of sea-level variations in modulating the depositional record in locations. An important recent finding of an almost complete sapropel
the different physiographic basin settings (platform, slope, abyssal series in shallow waters comes from the Adriatic Sea, where core
plain) that characterize siliciclastic systems, since these systems react Prad1-2 in approximately 70 m of sediments contains the last 370 ka
to sea-level oscillation by changing the balance between the rate of (Piva et al., 2008a,b) and the series of sapropels S1 to S10 except for S2.
sediment supply and the rate of creation of accommodation space. Because our data compilation (Table 1 and Fig. 1) shows that the
The sequence stratigraphic approach (sensu Van Wagoner et al., complete series of “Brunhes” sapropels is mostly recorded in the
1990) to study organic carbon-rich sedimentary intervals (sapropels and deeper sites (bathyal plain) of the basin but it is present in slopes and
black shales) documents their main occurrence within condensed shelves as well (Table 1), two questions arise: 1) what are the factors
sections that correspond to maximum flooding surfaces (Vail et al.,1984; that allow the presence of sapropels in the sediment record? 2) Does
Posamentier and Allen, 1999), or during the subsequent period of high the sapropel record depend on a low sedimentation rate and/or on a
sea-level, within 3rd-order cycles. Generally, the organic matter that different preservation potential?
accumulates under these conditions, has a marine origin (Type II An interesting paper by Murat and Got (2000) examines the TOC
kerogen, e.g. Klemme and Ulmisheck, 1991). On the other hand, very content in sapropel S1 occurrences to show that a positive correlation
important petroleum systems pertain to sediment sequences inter- exists between TOC concentration and depth and that, regardless of
preted as beeing deposited during periods of low sea-level (Henz and location or setting, the same percentage of organic carbon is
Zeng, 2003). In these cases accumulation of organic carbon-rich layers preserved at the same depth. As a consequence, the greater is the
mainly shows a high percentage of kerogen type III that is related to an organic carbon content the greater is the possibility to preserve the
increase of terrigenous supply (Klemme and Ulmisheck, 1991). sapropel record.
R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257 251

Table 1
Core data set used for this work

Lat. Long. Depth Thickness References


Western Mediterranean sites
ODP 650 Thyrrenian Sea Marsili Basin 39°21,40'N 13°54,05'E 3516.30 633.80 NO Kastens, Mascle et al., 1987.
ODP 651 Thyrrenian Vavilov Basin 40°09,03'N 12°45,39'E 3578.00 550.90 NO Kastens, Mascle et al., 1987.
ODP 652 Thyrrenian Sardinia continental 40°21,30'N 12°08,59'E 3446.00 721.10 NO BRUNHES SAPROPELS Kastens, Mascle et al., 1987.
Margin
ODP 653 Thyrrenian Cornaglia Terrace 40°15,86'N 11°26,98'N 2817.00 240.70 NO EVIDENT SAPROPEL Kastens, Mascle et al., 1987.
ODP 654 Thyrrenian Sardinia continental 40°34,76'N 10°41,80'E 2208.00 483.40 first sapropel 44 m depth Kastens, Mascle et al., 1987.
Margin
ODP 655 Thyrrhenian rise in the Vavilov basin 40°10,33'N 12°27,92'E 3290.00 90.40 First sapropel 8 m depth Kastens, Mascle et al., 1987.
ODP 656 Thyrrhenian Sea Monte de Marchi 40°11,96'N 12°11,03'E 3597.00 236.40 NO Kastens, Mascle et al., 1987.
ODP 963 Sicily Channel 37°01,938'N 13°10,896'E 481.00 199.40 NO Kastens, Mascle et al., 1987.
ODP 972 Ionian Sea Western Mediterranean 35°46,797'N 18°43,515'E 3942.00 95.40 NO Comas, Zahn et al., 1996.
Ridge
ODP 974 Thyrrhenian Sardinia Continental 40°21,364'N 12°08,506'E 3458.80 204.50 First sapropel 22 m depth Comas, Zahn et al., 1996.
Margin = ODP site 652
ODP 975 Balearic Sea Basin NW Flank 38°53,795'N 4°30,587'E 2427.60 317.10 First sapropel 3.8 m depth Comas, Zahn et al., 1996.
ODP 976 Alboran Basement Ridge 36°12,318'N 4°18,800'W 1107.50 928.70 NO SAPROPEL but continuous Comas, Zahn et al., 1996.
succession MIS 1–5 Cita et al. (1977), Capotondi
and Vigliotti, 1999,
von Grafenstein et al., 1999)
ODP 977 Alboran Basin 36°01,907'N 1°57,319'W 1995.50 598.50 first sapropel 2.0 m depth Comas, Zahn et al., 1996.
ODP 978 Alboran Basin 36°13,867'N 2°03,424'W 1941.00 698.00 NO Comas, Zahn et al., 1996.
ODP 979 Alboran Basement Ridge 35°43,427'N 3°12,353'W 1074.00 580.90 NO Comas, Zahn et al., 1996.
KET 8022 Thyrrhenian Sea 40°35,00 N 11°42,50' E 2430.00 No sapropel but all MIS (1–5) Paterne et al., 1986, 1988

Eastern Mediterranean complete succession


PRAD1-2 Adriatic Sea 42°40'34,7826qN 14°46'13,5565qE 185.50 71.20 S1, S3, S4, S5, S6, S7, S8, S9, S10 Piva et al., 2008a,b
KC01B Ionian Calabrian Ridge 36°15,25'N 17°44,34E 3643.00 S1, S3, S4, S5, S6, S7, S8, S9, S10 Sanvoisin et al., 1993;
Castradori, 1993
BD2002-GC01 Calabrian Rise 37°33.89671’N 17°48.32200’E 2470 3.41 S1, S3,S4,S5 Giunta et al. 2006
M25/4-13 Ionian slope 37°33.21'N 17°49.44' E 2533.00 16.30 S1, S3, S4, S5, S6, S7, S8, S9, S10 Schmiedl et al., 1998
M25/4-12 Ionian slope 37°57.98′N 18°11.04′E 2468.00 15.60 S1, S3, S4, S5, S6, S7, S8, S9, S10 Negri et al., 1999
D03 PC Mediterranean Ridge 34°53.16′N 20°48.54′E 2825.00 8.40 S1, S3, S4, S5 (2m), S6, S7, S8 Fusi et al., 1996
M40/1-22 Olimpi area 33°39.5′N, 24°41.2′E, 2004.00 7.00 S1, S2, S3, S4, S5 , S6, S7, S8, S9, Lowemark et al., 2006
S10
971 A Levantine Sea Napoli Volcano 33°42,190'N 24°42,814'E 2037.50 203.50 S1, S3?, S4?, S5?S6?, S7?, S8? Emeis, Robertson et al., 1996
RC9-181 South Crete Levantine basin 33°25,00'N 25°01,00E 2286.00 S1, S2, S3, S4, S5 , S6, S7, S8, Cita et al., 1977, Vergnaud-
(hiatus S3) Grazzini et al., 1977
KL51 ESE Crete 34°48.83’N 27°17.77’E 2158.00 7.00 S1, S3, S4, S5 , S6, S7, Frydas and Hemleben 2007
967 Levantino Base N flank Eratostene 34°04,098'N 32°43,523'E 2564.20 600.30 S1, S2, S3 , S4, S5 S6?, very Emeis, Robertson et al., 1996
bioturbated

Eastern Mediterranean missing S3 or S4 or both


SIN97-GC05 Medina Rise 34° 55’38”N 17°04’04”E 1075 4.52 S5, S6, S7, S8 Giunta et al., 2006
ODP 964 Ionian Calabrian Ridge 36°15,623'N 17°44,990'E 3668.00 101.80 S1, S5 Emeis, Robertson et al., 1996
ODP 973 Ionian Western Mediterranean 35°46,820'N 18°56,889'E 3706.50 148.50 S1, S?, S5, S6, S7? Emeis, Robertson et al., 1996
Ridge
M40/1-10 Mediterranean Ridge 34°45.8′N 19°45.7′E 2972.00 8.3 S1, S5 , S6, S7, S8, (hiatus S3) Lowemark et al., 2006
D05 PC Mediterranean Ridge 34° 52.64′N 20° 48.80′E 2817.00 5.38 S1, S4, S5, S6, Fusi et al., 1996
D04 PC Mediterranean Ridge 34° 52.80′N 20°51.27′E 2680.00 8.26 S1,S?, S5, Fusi et al., 1996
D10 PC Mediterranean Ridge 34° 52.78′N 20°46.78′E 2671.00 5.66 S1, S?, S6, S8 Paterne et al., 1986, 1988
U08 GC Mediterranean Ridge 34° 59'26.40′N 21°01' 08.09qE 2150.00 5.88 S1, S4, S5, S6, S7 Fusi et al., 1996
UM94PC17 Urania basin Area 35°13’56.4”N 21°24’42.6”E 3224 5.00 S1,S5,S6 Giunta et al., 2006
U01 GC Mediterranean Ridge 35°11'48qN, 21°24' 45qE 3227.00 3.75 S1, S5, S6, S7, S8 Fusi et al., 1996
SIN97-GC01 SW Crete 35° 07’27’’N 22°42’02.31”E 933 3.08 S1, S4,S5 Giunta et al., 2006
BAN84-GC27 Katia basin 33° 20.99’N 23°13.26’E 2026 4.96 S1, S5, S6,S7, S8, S9 Giunta et al., 2006
U06 PC Mediterranean Ridge 35°07'15,21qN 24°15' 05.43′E 2582.00 4.72 S1, S4, S5, S6, S7, S8 Fusi et al., 1996
BAN88-GC11 Area Prometheus 2 33° 49.10’N 24°26’E 1910 5.62 S1, S4,S5, S6,S7,S8 EEMIANO Project

Eastern Mediterranean uncertainty in missing sapropels


BAN89-GC09 Napoli dome 33° 37.98’N 24°43.29’E 2011 3.27 S1, S4,S5 Giunta et al., 2006
ODP 969 Levantine-Mediterranean Ridge 33°50,399'N 24°53,066'E 2211.80 108.30 S1,S3,S5, S6,S7,S8 Emeis, Robertson et al., 1996
BAN82-GC16 Erodothus Abyssal Plain 32° 34.42’N 26°50.85’E 3008 4.47 S1,S5,S6, S8 Giunta et al., 2006
ODP 966 Levantine Eratostene Top 33°47,799'N 32°42,095'E 937.70 106.80 S1, S5? S6, S?, S? Emeis, Robertson et al., 1996
ODP 965 Levantine Eratostene Northern flank 33°55,080'N 32°42,785'E 1517.70 250.40 Only S1 Emeis, Robertson et al., 1996
ODP 970 Levantino Volcano Milano 33°44,194'N 24°48,120'E 2086.90 201.40 Only S1(short core) Emeis, Robertson et al., 1996

An important factor that can affect the sapropel record is top-down This work indicates that the original thickness and TOC concentration
oxidation as reported by Löwemark et al. (2006). In fact, the so called influence the burn-down rate and that local and regional setting, in
sapropel burn-down is a process implying that sapropels can be relation to deep water circulation, plays a major role in the
partially and sometimes completely oxidized. Löwemark et al. (2006) preservation of sapropels even if variations in the sedimentation
show that all sapropels from S1 to S11 experienced variable top-down rate are evident. In particular, cores collected far from well oxygenated
oxidation and that none of them maintained their original thickness. watermasses generally display better preserved sapropels.
252 R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257

Fig. 1. Eastern Mediterranean Sea map showing the location of the cores listed in Table 1. Black stars refer to cores in which a complete sapropel S5 to S1 series has been described.
White circles refer to cores where the sapropels S5 to S1 series is not complete. The white cross shows the location of Chirp sonar AMC 237 and AMC 178 (see Ridente and Trincardi
2002 and 2005 for details).

Finally, an important conclusion of Löwemark et al. (2006) is that part of this latter interval, spanning from about 35 ka to 8 ka, record
the best preserved sapropels exist when a succession of clustered the sea-level fall that continued until about 12 ka (Ridente and
strong insolation maxima occurred. In contrast, moderate or weak Trincardi, 2005). In Table 2 we report the values that we obtained from
insolation maxima followed by weak insolation maxima are asso- three cores, each representative of a different depositional condition
ciated with the poorest preserved sapropels, often only detectable as (shelf, lower slope, bathyal plain). Major findings are: 1) the mean
“ghost” sapropels through geochemical analysis. sedimentation rate in the bathyal plain and in the lower slope remains
Therefore, the data suggest that although sapropels were depos- in the same order of magnitude, even if it decreases during the interval
ited throughout the eastern Mediterranean and Adriatic sea, their Y5-S1. This lower sedimentation rate can couple with African dryer
sedimentary record can be affected by strong burn-down effects. climate during which the entire Nile system must have been reduced
to intermittent seasonal flow (Lamb et al., 2007 cum bibl.); 2) the
3. Stratigraphic and sedimentary framework sedimentation rate between S5 and Y5 on the Adriatic shelf is one
order of magnitude higher than at deeper locations and, between Y5
In order to highlight possible variations in sediment supply and and S1 a further increase of terrigenous supply was directly linked to
their relations with variations in sea-level during the studied time the sea-level fall and the developing forced-regressive wedge.
interval, we examined published data and calculated sedimentation The sedimentary sequence on the shelf is, obviously, the most
rates in different depositional systems, such as shelf, slope and bathyal appropriate to obtain information on the relation between the relative
plain (Table 2). We used as an age marker the Y5 tephra (Keller et al., position of sea-level and the deposition of sapropelic intervals. In fact,
1978; Emeis et al., 1996), which is the closest layer to the Last Glacial this depositional system is directly influenced by variations of the
Maximum (LGM). We calculated the average sedimentation rate terrigenous supply induced by sea-level oscillations. For this reason,
between insolation-cycle 12 (S5 - Hilgen, 1991)) and Y5. Then, we we referred to the abundant literature regarding the Adriatic Sea in
calculated the average sedimentation rate between Y5 and the order to locate the position of different sapropels within the
subsequent transgression up to insolation-cycle 2 (S1). The main sedimentary sequences occurring in this area. In particular, according
to Ridente and Trincardi (2005, cum bibl.), the upper Pleistocene–
Holocene stratigraphic succession of the central portion of the western
Table 2
Sedimentation rate in the eastern Mediterranean and Adriatic sea calculated using S1, Adriatic margin records relative sea-level variations related to ca.
S5 and Y5 data point (see text) 100 ka glacio-eustatic cycles. Following the Last Glacial Maximum and
until ca. 6 ka BP, sea-level rose about 125 m, or up to 149 m (Lambeck et
Core Location Depth S1–Y5 sed. rate Y5–S5 sed. rate
(mm/year) (mm/year) al., 2004; Antonioli and Vai, 2004) from its lowstand position to the
967 Levantino Base 2564.20 0.024 0.070
modern highstand (Fairbanks, 1989; Bard et al., 1996). According to
N flank Eratostene Ridente and Trincardi (2005), four depositional sequences (numbered
973 Ionian Western 3706.50 0.012 0.043 1 to 4, topdown) bounded by regional erosional unconformities and
Mediterranean Ridge their distal conformities (namely ES1 to ES5) record cycles older than
PRAD1-2 Adriatic Sea 185.50 0.38 0.200
the Holocene (between ca. 25 and 450 ka). The four Pleistocene
R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257 253

depositional sequences appear internally composed of laterally Trincardi, 2002). Sapropels S5 to S1 have been located on the
continuous forced-regressive units, extending from the western flank chronostratigraphic chart based on correlation with the sea-level
of the Middle Adriatic Depression (MAD) to the area south of the curve and Marine Isotopic Stages (Löwemark et al. 2006; Martinson
Gargano Promontory. The uppermost forced-regressive unit (sequence et al.,1987). It can be observed that sapropels S5 and S1 are located in the
1 of Ridente and Trincardi, 2005) is the result of the Po delta same depositional context (TST), whereas sapropels S4, S3 and S2 fall
progradation and is younger than 30–35 ka (Ridente and Trincardi within sequence 1 at different stratigraphic levels, within the Highstand
2005) based upon the presence of the Y5 “Campanian Ignimbrite” Systems Tract or, in the case of S2, within the Falling-stage Systems Tract
tephra layer (35 ka n.c. age, Calanchi et al., 1996, 1998) close to its base. (FST). Sapropels S4, S3 and S2, in Fig. 2, also correlates with the TST of a
Dating of the sequences based on the analysis of several cores 5th-order sequence within a HST and an FST of a 4th-order sequence.
(Trincardi et al., 1996; Asioli, 1996) allows reference of erosional In addition, we examined two acoustic Chirp sonar profiles: AMC 237
surface ES1, which truncates sequence 1, to the sea-level fall and (Ridente and Trincardi 2005), located south east of the MAD, oriented
lowstand of the Last Glacial Maximum (Oxygen Isotope Stage 2, ca. 20– SW-NE, and AMC 178 (Ridente and Trincardi, 2002) located north of the
30 ka) and identifies deposits from Stages 4 and 5 within the lower part Gargano promontory and oriented SE-NW that are representative of the
of sequence 1 (Ridente and Trincardi 2005). The underlying erosional sedimentary record (Ridente and Trincardi, 2002).
surface ES2 was indirectly referenced to the lowstand of Isotope Stage On the line-drawing of these profiles (Fig. 3) we located the ES1
6 (ca. 130–140 ka, after Martinson et al., 1987; Ridente and Trincardi, and ES2 sequence boundaries (as indicated by Ridente and Trincardi,
2005). 2002, 2005) and the geometry of the major reflectors.
Above the ES1 sequence boundary, the late Pleistocene–Holocene According to Fig. 2, therefore, sapropel S5, on the deeper north-
Trangressive Systems Tract (TST) and the late Holocene Highstand eastern part of the profile AMC 237, should be located in the lower part of
Systems Tract (HST) show shore-parallel depocentres (Cattaneo and Sequence 1, within a plane-parallel unit that represents an aggradational
Trincardi, 1999; Correggiari et al., 2001), likely suggesting advection of phase of deposition that Ridente and Trincardi interpreted as the TST. S4
fine-grained sediment as the shelf was progressively drowned during and S3, in contrast, should occur in the upper part of the sequence,
the sea-level rise (Cattaneo and Trincardi, 1999). Muddy deposits, which shows low-dipping-angles oblique clinoforms. Ridente and
more than 30 m thick (Ciabatti et al., 1987), characterize the highstand Trincardi (2002) suggest that this geometry records a condition of
interval that accumulated on the shelf during the last ca. 6 ka enhanced sediment supply on the inner shelf during the late highstand
(Holocene HST, Correggiari et al., 2001; Cattaneo et al., 2003). and early falling of sea-level. In the AMC 178 the HST is well developed
In the chronostratigraphic chart of Fig. 2 the systems tracts have been and the overlying unit shows higher dipping angles and evident
drawn by correlation between sea-level position (after Waelbroeck et al., downlap terminations, linked to the deposition of the regressive
2002) and the geometry and distribution of the seismic facies in the wedge during the subsequent sea-level fall (Ridente and Trincardi,
central portion of the western Adriatic margin (cfr Fig. 3, Ridente and 2002). As the base of the distal forced-regressive wedge is dated to after

Fig. 2. Chronostratigraphic chart which displays both the horizontal distribution of the component sedimentary layers of the studied sequences and the significant hiatuses in
sedimentation. The systems tracts have been drawn by means of the correlation between the sea level position (after Waelbroeck et al., 2002) and the geometry and distribution of
the seismic facies in the central portion of the western Adriatic margin (cfr Fig. 3 Ridente and Trincardi, 2002). The position of sapropels S5 to S1 has been reported on the
chronostratigraphic scheme after the correlation of their occurrence with respect to the sea level curve and Marine Isotopic Stages (Löwemark et al 2006; Martinson et al., 1987).
254 R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257

Fig. 3. Line drawing of Chirp sonar AMC 237 and AMC 178 (see Ridente and Trincardi 2005 and 2002 for details). Definition of the different system tracts derives from the geometry of
the seismic reflectors according to Ridente and Trincardi (2002, 2005)) see text for further explanation.

30–35 ka, this means that in this sequence we should also observe condensed interval commonly includes OC-rich layers composed of
sapropel S2 (55 ka). Its occurrence, however, falls within the well- mostly pelagic and hemipelagic clastics that are widely distributed
developed prograding clinoforms, linked to enhanced sediment supply throughout the basin (e.g. Vail et al., 1984; Wignall, 1994). At this time,
during the late highstand and falling stage of sea-level. For this reason pelagic sedimentation is less affected by clastics and is more sensitive to
terrigenous sediment supply likely prevented formation of an S2 Corg- the paleoceanographic setting, allowing a better recording of enhanced
rich horizon in this framework. primary productivity and/or density stratification in the water column.
Above the ES1 sequence boundary, the Holocene succession shows a Furthermore, under these conditions, the Particulate Organic Matter
plane-parallel geometry of the seismic reflectors. The Holocene mud (POM) settling velocity increases due to entrapment of pelites in marine
deposits can be divided into a lower unit TST, which includes the S1 snow and to the production of large fecal pellets. This fact partially
horizon, and an overlying HST the base of which, according to Trincardi prevents organic matter recycling in the water column, contributing to
et al. (1996) and Cattaneo and Trincardi (1999), is dated at 5.5 ka in the sedimentation of a larger amount of organic carbon (Stow et al., 2001;
studied region. Therefore, in the Holocene sequence the stratigraphic Bianchi et al., 2006). These processes proceed until the prograding shelf-
position of S1 is equivalent to that of S5 in the underlying sequence 1, wedge during the HST (progradational stacking patterns of Van
showing the same depositional context as most sapropels and black Wagoner et al., 1990) overwhelm the marine organic carbon (OC)
shales described in the literature. According to Van Wagoner et al. (1988, input. At this latter time, OC-rich intervals can be deposited where low
1990), successive parasequences are built as sea-level rises, generating a sedimentation rates and muddy deposits persist up to the late HST. This
retrogradational stacking pattern (TST). As sea-level reaches highstand occurs in the slope, where hemipelagic deposition includes vertical
conditions, the shelf is under fairly deep water and is starved for settling through the water column and slow lateral advection (Stow et
sediment. A condensed section, a thin marine muddy interval, al., 2001) and in the basin plain, where pelagic settling is only
characterized by very low depositional rate, is deposited. This episodically interrupted by turbidity currents.
R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257 255

An example is provided by Capozzi et al. (2006) regarding general, overall deep-water formation, causing a slow down or even
sapropels deposited during the Middle Pliocene HST (3rd-order stoppage of thermohaline ventilation in the deeper parts of the
cycle) pertaining to the Cluster “O” (Verhallen, 1987; Lourens et al., Mediterranean, favouring the formation of anoxic bottom water. At the
1996). The Pliocene 3rd order cycle, spanning from 3.75 Ma to 2.87 Ma same time, the increased runoff combined with a shallowing of the
(Capozzi and Picotti, 2003), records a sea-level drop (LST) and a nutricline into the photic zone led to enhanced primary productivity and
widespread turbiditic deposition, interpreted to be related to an therefore an increased export of organic material to the sea floor
important glacial event in the Northern Hemisphere. After this event, (Rohling and Gieskes, 1989; Rohling, 1991, 1994, 2001; Corselli et al.,
a climatic optimum (TST and early HST) from 3.3 to 3.042 Ma was 2002). It is worth noting in this framework the modelling results
marked by the growth of a carbonate platform. A climatic deteriora- obtained by Bianchi et al. (2006), who pointed that sapropel S1
tion, at the end of the warm Middle Pliocene between 3.042 and deposition occurred by the concomitant effects of absent reventilation
2.87 Ma, occurred during a period of stable sea-level (late HST) and and enhanced productivity with the additional contribution of a large
caused the demise of the carbonate platform and the deposition of POM sinking velocity.
sapropel cycles. These cycles record a constant sedimentation rate, This reconstruction fits also for the sapropel S5 which was deposited at
which indicates that variations in runoff and, consequently, in nutrient the beginning of an interglacial and therefore during a rapid sea-level rise
supply, did not occur in the depositional system, whereas orbital coupled with increased runoff. At times of deposition of sapropels S4 and
increases in the intensity of solar radiation were the main reason for S3, the sea-level position, as reconstructed after Waelbroeck, 2002 (Fig. 2),
enhanced primary productivity. The deposition of these sapropel was −25 m lower than today, while sapropel S2 was deposited when the
layers occurred on the upper slope, where mud deposition continued sea-level was at least 50 m lower than today. The lowered sea-level means
until a rapid fall in sea-level changed the sedimentary processes in that the northern Adriatic coastline was shifted few kilometres southward
this location. This situation lasted about 140 ka during the Mid- during deposition of S3 and S4, and more that 100 km in the case of S2.
Pliocene, but it is easily comparable to the sedimentary conditions Cold, deep waters form in the Adriatic Sea during the winter in
described for S4 and S3. Furthermore, the deposition of the upper response to heat loss to the atmosphere (Artegiani et al., 1989, 1997).
Sapropel E (in Capozzi et al., 2006) may account for the sedimentary In particular, the North Adriatic Deep Water (NADW) is a water mass
conditions experienced during deposition of the S2 interval. In the formed in the north Adriatic Sea during strong “Bora” wind events.
case of Sapropel E, in fact, the increase of sand supply not only strongly The very dense deep water flows southward along isobaths near the
diluted the marine organic carbon but also contributed terrestrial bottom of the Italian shelf. The flow partially sinks into the MAD, but
organic carbon. its major portion moves further southwards to reach the shelf off Bari,
During sea-level fall, the shallower portions of the sea floor of epeiric where a canyon intersects the shelf and the water deepens. At this
seas and the continental shelf are progressively exposed to subaerial point the NADW plays a role in the formation of Adriatic Bottom Water
erosion. The erosion causes formation of the sequence boundaries, which (ABW) (see Artegiani et al., 1993).
are diachronous and cap the previous HST, also eroding the surface of the During sedimentation of sapropel S2, extensive shelf areas had
downstepping sediments, deposited during forced regression (FST) that is emerged in the north of the basin, implying also a reduction in the size of
associated with the sea-level fall (Catuneanu, 2002). the water-mass subject to the “Bora” winds. According to the data shown
When lowstand is attained (LST), sediments transported down in Russo and Artegiani (1996), the lowest winter water temperatures are
valleys are delivered mainly off the shelf edge into deep water, to form reached in the northern Adriatic between the Northwestern coastline and
sea floor fans or lowstand wedges. Downslope resedimentation approximately the 20 m isobath. Therefore, it is probable that, at the time
processes, such as turbidity currents, debris flow and slides, are of sapropel S2 formation, the NADW was weakened because large shelf
dominant and the sedimentation rate increases in the basin plain. areas were exposed and as a consequence the formation of the ABW was
These sediments can accumulate high amounts of organic carbon and its diminished. An implication of these factors is that the conditions for
rapid burial prevents its oxidation (Stow et al., 2001). In this lowstand sedimentation of sapropel S2 should have been facilitated because of the
scenario, however, the occurrence of mainly gravity- and density-driven slow down of thermohaline circulation and decreased ventilation at the
sedimentary processes produce high-frequency oxygen replenishment sea bottom. However, sapropel S2 is rarely found and its expression
at the basin-floor and also cause dilution of primary-productivity corresponds to a not well expanded insolation maximum (Hilgen, 1991)
derived organic matter. Organic matter accumulation in sapropelic suggesting that the primary productivity did not increase sufficiently to
layers can to some extent survive in the slope environment, where lutite cause enough organic matter accumulation. This observation strengthen
flow and cascading processes govern the deposition of fine-grained the hypothesis that productivity played a major role in the formation of
sediment, during cyclic episodes of enhanced runoff. sapropels and, for this reason, the S2 is a not well developed sapropel. This
Summarizing, the progradation of the shelf-wedge progressively situation makes the sapropel S2 more suitable for burn-down, and, as
shifts sapropel deposition basin-ward that can be expressed from the demonstrated by Löwemark et al. (2006), the fate of sapropel S2 was to be
basin margin to the deep basin-plain only when terrigenous sediment erased in many cases, eliminating its appearance in the vast majority of
starvation is reached. During times of low sea-level, the slope and the the sediment record far collected at the sea bottom.
basin plain, if not reached by turbidity currents, are environments that
can better record enhanced organic carbon accumulation during 5. Conclusions
higher-frequency sea-level oscillations.
Sapropels levels S5 to S1, clustered in high frequency Milankovian
4. Sea-level control on watermass circulation cycles, occur within a sequence stratigraphic framework consisting of
4th order cycles linked to late Pleistocene–Holocene sea-level
According to the data discussed in this paper, it is clear that oscillations, which have amplitudes of over 120 m and have been
deposition of sapropels occurred under different sea-level conditions. correlated to glacial-interglacial cycles modulated by the 100 ka
Whereas sapropels S5 and S1 were deposited during a rapid sea-level eccentricity.
rise, related to the deglacial phase, the others (S4, S3, S2) were deposited
during an interval recording sea-level still-stand and subsequent sea- 1) In terms of sequence stratigraphic interpretation, sapropels S5 and
level fall. S1 in the central Adriatic shelf and MAD occur in the Transgressive
According to the most recent models regarding sapropel S1 Systems Tract of two depositional sequences, that are assigned to MIS
formation, increased fresh water input lowered the salinity of the 5e and 1, respectively. Sapropels S4 and S3 are found within the HST
surface waters and stopped the formation of Deep Adriatic Water and, in of late Pleistocene sequence 1 (Ridente and Trincardi, 2002, 2005),
256 R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257

which developed during MIS 5. Sapropel S2, in contrast, corresponds eastern tropical Atlantic at orbital time scales (ODP Site 959, off Ivory Coast. In:
Harris, N., Pradier, B. (Eds.), The Deposition of Organic-Carbon-Rich Sediments:
to the warm MIS 3.3 and is part of the FST. It is noteworthy that S4, S3 Models, Mechanisms, and Consequences, SEPM Spec. Pub., 82, pp. 125–143.
and S2 were deposited during the TST of a 5th-order sequence within Bianchi, D., Zavatarelli, M., Pinardi, N., Capozzi, R., Capotondi, L., Corselli, C., Masina, S.,
2006. Simulation of ecosystem response during the sapropel S1 deposition event.
a HST and an FST of a 4th-order sequence.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 235, 265–287.
2) Sapropels S5 and S1 are widespread from the basin margins to the Calanchi, N., Dinelli, E., Lucchini, F., Mordenti, A., 1996. Chemostratigraphy of late
plains of the Eastern Mediterranean. Under conditions of sea-level Quaternary sediments from lake Albano and Central Adriatic sea cores (PALICLAS
Project). In: Guilizzoni, P., Oldfield, F.L. (Eds.), Palaeoenvironmental Analysis of
rise, the sedimentary basin experienced starvation of terrigenous Italian Crater Lake and Adriatic Sediments (PALICLAS progect). Mem. Istit. It. Idrob.,
sediments and therefore was more favourable to sapropel deposition. 55, pp. 247–263.
Calanchi, N., Cattaneo, A., Dinelli, E., Gasparotto, G., Lucchini, F., 1998. Tephra layers in
Comparison of the S5 to S2 sapropel series with Mid-Pliocene Late Quaternary sediments of the Central AdriaticSea. Mar. Geol. 149, 191–209.
cluster “O” shows that, in both cases, sapropels may form until the Calvert, S.E., 1983. Geochemistry of Pleistocene sapropel and associated sediments from
the Eastern Mediterranean. Oceanol. Acta 6, 255–267.
prograding shelf wedge of the late highstand and initial falling Capotondi, L., Vigliotti, L.,1999. Magnetic and microfaunal characterization of late Quaternary
stage in sea-level provides an increase in sediment supply, leading sediments from the western Mediterranean: inferences about sapropel formation and
paleoceanographic implications. In: Zahn, R., Comas, M.C., Klaus, A. (Eds.), Proc. ODP, Sci.
to progressive dilution of marine organic carbon. This conditions
Results,161: College Station, TX (Ocean Drilling Program), pp. 505–518. doi:10.2973/odp.
led to formation of sapropel S2 and the uppermost sapropel of proc.sr.161.243.1999.
cluster “O” (Capozzi et al., 2006). Capozzi, R., Picotti, V., 2003. Pliocene sequence stratigraphy, climatic trends and
sapropel formation in the Northern Apennines (Italy). Palaeogeogr. Palaeoclimat.
3) When lowstand conditions are attained, marine sedimentation Palaeoecol. 190, 349–371.
shifts off the shelf break, via mainly gravity-driven processes. In this Capozzi, R., Dinelli, E., Negri, A., Picotti, V., 2006. Productivity-generated annual laminae
in Mid-Pliocene sapropels deposited during precessionally forced periods of
case, the increase in terrigenous supply in deeper areas may
warmer Mediterranean climate. Palaeogeogr. Palaeoclimat. Palaeoecol. 235,
contribute to storage of high amounts of organic carbon, although 208–222.
dilution and mixing with terrestrial organic carbon commonly occur. Castradori, D., 1993. Calcareous nannofossils and the origin of eastern Mediterranean
sapropels. Paleoceanograpy 8, 459–471.
4) Low sea-level affects circulation in the Adriatic Sea and is likely Cattaneo, A., Trincardi, F., 1999. The late-quaternary transgressive record in the Adriatic
to cause a consequent weakening of formation of North Adriatic Epicontinental Sea: basin widening and facies partitioning. In: Bergman, K.,
Snedden, J. (Eds.), Isolated Shallow Marine Sand Bodies: Sequence Stratigraphic
Deep Water. This condition, which could prevent normal sea floor Analysis and Sedimentologic Interpretation, SEPM, Spec. Publ., 64, pp. 127–146.
ventilation, however, does not correspond to a well developed Cattaneo, A., Correggiari, A., Langone, L., Trincardi, F., 2003. The late-Holocene Gargano
subaqueous delta, Adriatic shelf: sediment pathways and supply fluctuations. Mar.
sapropel S2. Productivity increases, paced by well developed Geol. 193, 61–91.
precession minima (insolation maxima), are suggested to play a Catuneanu, O., 2002. Sequence stratigraphy of clastic systems: concepts, merits, and
pitfalls. J. Afr. Earth Sci. 35 (1), 1–43.
major role in sapropel deposition. Therefore, we suggest that
Ciabatti, M., Curzi, P.V., Ricci Lucchi, F., 1987. Quaternary sedimentation in the Central
sapropel S2 is virtually missing in the eastern Mediterranean basin Adriati Sea. Giorn. Geol., ser. 3° 49 (1), 113–125.
because of a low amplitude insolation maximum, and consequent Cita, M.B., Vergnaud-Grazzini, C., Robert, C., Chamley, H., Ciaranfi, N., D’ Onofrio, S., 1977.
Paleoclimatic record of a long deep sea core from the eastern Mediterranean. Quat.
insufficient productivity increase, combined with post-deposi- Res. 8, 205–235.
tional oxidation that erased its record. Cobianchi, M., Picotti, P., 2001. Sedimentary and biological response to sea-level
and palaeoceanographic changes of a Lower–Middle Jurassic Tethyan platform margin
Acknowledgments (Southern Alps, Italy). Palaeogeogr. Palaeoclimat. Palaeoecol. 169 (3-4), 219–244.
Comas, M.C., Zahn, R., Klaus, A., et al., 1996. Proc. ODP, Init. Repts., 161: College Station,
TX (Ocean Drilling Program). doi:10.2973/odp.proc.ir.161.1996.
The authors are indebted to Giambattista Vai and Vincenzo Picotti Correggiari, A., Trincardi, F., Langone, L., Roveri, M., 2001. Styles of failure in late Holocene
for critically reading the, first version of the manuscript. We warmly highstand prodelta wedges on the Adriatic shelf. J. Sed. Res. 71 (2), 218–236.
Corselli, C., Principato, M.S., Maffioli, P., Crudeli, D., 2002. Changes in planktonic
thank Alessandro Amorosi and an anonymous reviewer whose assemblages during sapropel S5 deposition: evidence from Urania Basin area,
comments greatly help to improve the paper. RFO Funding to R.C. is eastern Mediterranean. Paleoceanography 17 (3), 1029. doi:10.1029/2000PA000536.
gratefully acknowledged. Cramp, A., O’Sullivan, G., 1999. Neogene sapropels in the Mediterranean: a review. Mar.
Geol. 153, 11–28.
EEMIANO Project, http://www.geo.unimib.it/Eemianoweb/Cores_S5.htm.
References Emeis, K.C., Shipboard Scientific Party of Leg 160, 1996. Paleoceanography and sapropel
introduction. In: Emeis, K.C., Robertson, A.H.F., Richter, C., et al. (Eds.), Proc. ODP Init.
Aksu, A.E., Ya,sar, D., Mudie, P.J., 1995. Paleoclimatic and paleoceanographic circum- Repts. College station TX (Ocean Drilling Program), 160, pp. 21–27.
stances leading to the development of sapropel layer S1 in the Aegean Sea basins. Emeis, K.-C., Sakamoto, T., Wehausen, R., Brumsack, H.-J., 2000a. The sapropel record of
Palaeogeography Palaeoclimatology Palaeoecology 116 71–101. the eastern Mediterranean Sea—results of Ocean Drilling Program Leg 160.
Antonioli, F., Vai, G.B., 2004. Litho-paleoenvironmental maps of Italy during the last two Palaeogeogr. Palaeoclimatol. Palaeoecol. 158, 371–395.
climatic extremes. Explanatory Notes. 32nd International Geological Congress, Emeis, K.-C., Struck, U., Schulz, H.-M., Rosenberg, R., Bernasconi, S., Erlenkeuser, H.,
Firenze. Sakamoto, T., Martinez-Ruiz, F., 2000b. Temperature and salinity variations of
Arthur, M., Sageman, B., 2005. Sea-level control on source-rock development: Mediterranean Sea surface waters over the last 16,000 years from records of
perspectives from the Holocene Black Sea, the mid-Cretaceous Western Interior planktonic stable oxygen isotopes and alkenone unsaturation ratios. Palaeogeogr.
Basin of North America, and the Late Devonian Appalachian Basin. In: Harris, N., Palaeoclimatol. Palaeoecol. 158, 259–280.
Pradier, B. (Eds.), The Deposition of Organic-Carbon-Rich Sediments: Models, Fairbanks, R.G., 1989. A 17,000-year glacio-eustatic sea level record–influence of glacial
Mechanisms, and Consequences, SEPM Spec. Pub, 82, pp. 35–59. melting rates on the Younger Dryas event and deep-sea circulation. Nature 342,
Artegiani, A., Gacic, M., Michelato, A., Kovacevic, V., Russo, A., Paschini, E., Scarazzato, P., 637–642.
Smircic, A., 1993. The Adriatic Sea hydrography and circulation in spring and Frydas, D., Hemleben, C., 2007. Opal phytoplankton assemblages of the Late Quaternary
autumn (1985–1987). Deep. Sea Res. II 40, 1180. sapropel layers S5 and S7 from the southeastern Mediterranean Sea (“Meteor”-
Artegiani, A., Azzolini, R., Salusti, E., 1989. On dense water in the Adriatic Sea. Oceanol. Cruise 404. Site 67) Revue de Micropaleontologie, 50(2), pp. 169–184.
Acta 12 (2), 151–160. Fusi, N., Aloisi de Larderel, G., Borello, A., Amelio, O., Castradori, D., Negri, A., Rimoldi, B.,
Artegiani, A., Bregant, D., Paschini, E., Pinardi, N., Raicich, F., Russo, A., 1997. The Adriatic Sanvoisin, R., Tarbini, P., Cita, M.B., 1996. Marine geology of the Medriff Corridor,
sea General Circulation. Part I: air–sea interactions and water mass structure. Mediterranean Ridge. The Island Arc 5, 420–439.
J. Phys. Oceanography 27, 1492–1514. Giunta, S., Negri, A., Maffioli, P., Sangiorgi, F., Capotondi, L., Morigi, C., Principato, M.S.,
Asioli, A., 1996. High resolution foraminifera biostratigraphy in the Central Adriatic Corselli, C., 2006. Phytoplankton Dynamics in the Eastern Mediterranean Sea during
basin during the last deglaciation: a contribution to the PALICLAS Project. In: Marine Isotope Stage 5e. Palaeogeogr. Palaeoclimatol. Palaeoecol. 235, 28–47.
Old”eld, F., Guilizzoni, P. (Eds.), Palaeoenvironmental Analysis of Italian Crater Lake Hassold, N., Rea, D.K., Meyers, P.A., 2003. Grain-size evidence of the delivery mode of
and AdriaticSediments. Mem. Istit. It. Idrobiol., 55, pp. 197–218. terrigenous sediment components to a middle Pleistocene interrupted sapropel
Bard, E., Hamelin, B., Arnold, M., Montaggioni, L., Cabioch, G., Faure, G., Rougerie, F., from ODP Site 969, Mediterranean Ridge. Palaeogeogr. Palaeoclim. Palaeoecol. 190,
1996. Deglacial sea-level record from Tahiti corals and the timing of global 211–219.
meltwater discharge. Nature 382, 241–244. Henz, T.F., Zeng, H., 2003. High-frequency Miocene sequence stratigraphy, offshore
Beckmann, B., Wagner, T., Hofmann, P., 2005. Linking Coniacian–Santonian (OAE3) Louisiana: cycle framework and influence on production distribution in a mature
black-shale formation to African climate variability: a reference section from the shelf province. AAPG bull. 87, 197–230.
R. Capozzi, A. Negri / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 249–257 257

Hilgen, F.J., 1991. Astronomical calibration of Gauss to Matuyama sapropels in the Ridente, D., Trincardi, F., 2005. Pleistocene ‘‘muddy’’ forced regression deposits on the
Mediterranean and implication for the geomagnetic polarity time scale. Earth Adriatic shelf: a comparison with prodelta deposits of the late Holocene highstand
Planet. Sci. Lett. 104, 226–244. mud wedge. Mar.Geol. 222–223, 213–233.
Huc, A.Y., 1988. Sedimentology of organic matter. In: Frimmel, F.H., Christman, R.F. Rohling, E.J., 1991. Shoaling of the eastern mediterranean pycnocline due to reduction of
(Eds.), Humic substances and their role in the environment. Wiley, Chichester, excess evaporation: implications for sapropel formation. Paleoceanography 6,
pp. 215–243. 747–753.
Kastens, K.A., Mascle, J., Auroux, C., et al., 1987. Proc. ODP, Init. Repts., 107: College Rohling, E.J., 1994. Review and new aspects concerning the formation of eastern
Station, TX (Ocean Drilling Program). doi:10.2973/odp.proc.ir.107.1987. Mediterranean sapropels. Mar. Geol. 122, 1–28.
Keller, J., Ryan, W.B.F., Ninkovich, D., Altherr, R., 1978. Explosive volcanic activity in the Rohling, E.J., 2001. The dark secret of the Mediterranean - a case history in past
Mediterranean over the past 200,000 years as recorded in deepsea sediments. Bull. environmental reconstruction. http://www.soes.soton.ac.uk/staff/ejr/DarkMed/
Geol. Soc. Am. 89, 591–604. dark-title.html.
Kidd, R.B., Cita, M.B., Ryan, W.B.F., 1978. Stratigraphy of eastern Mediterranean sapropel Rohling, E.J., Gieskes, W.W.C., 1989. Late Quaternary changes in Mediterranean
sequences recovered during Leg 42A and their paleoenvironmental significance. intermediate water density and formation rate. Paleoceanography 4, 531–545.
Initial Reports of the Deep-Sea Drilling Project 42A, 421–443. Rossignol Strick, M., 1983. African monsoon an immediate climate response to orbital
Klemme, H.D., Ulmisheck, G.F., 1991. Effective petroleum source rocks of the world- insolation. Nature 304, 46–49.
stratigraphic distribution and controlling depositional factors: AAPG Bull, 75, Rossignol Strick, M., 1985. Mediterranean Quaternary sapropels, an immediate response
pp. 1809–1851. of the African monsoon to variations of insolation. Palaeogeogr. Palaeoclimatol.
Lamb, H.F., Bates, R.C., Coombes, P.V., Marshall, M.H., Umer, M., Davies, S.J., Dejen, E., Palaeoecol. 49, 237–263.
2007. Late Pleistocene desiccation of Lake Tana, source of the Blue Nile. Quat. Sci. Rossignol Strick, M., 1999. The Holocene climatic optimum and pollen record of
Rev. 26, 287–299. sapropel 1 in the eastern Mediterranean, 9000–6000 BP. Quat. Sci. Rev. 18, 515–530.
Lambeck, K., Antonioli, F., Purcell, A., Silenzi, S., 2004. Sea level change along the Italian Rossignol-Strick, M., Nestero¡, W., Olive, P., Vergnaud-Grazzini, C., 1982. After the
coast for the past 10,000 yrs. Quat. Sci. Rev. 23, 1567–1598. deluge: Mediterranean stagnation and sapropel formation. Nature 295, 105–110.
Lourens, L.J., Antonarakou, A., Hilgen, F.J., Van Hoof, A.A.M., Vergnaud-Grazzini, C., Roveri, M., Taviani, M., 2003. Calcarenite and sapropel deposition in the Mediterranean
Zachariasse, W.J., 1996. Evaluation of the Plio-Pleistocene astronomical timescale. Pliocene: shallow- and deep-water record of astronomically driven climatic events.
Paleoceanography 11 (4), 391–413. Terra Nova, 15, 279–286.
Löwemark, L., Lin, Y., Chen, H.-F., Yang, T.-N., Beier, C., Werner, F., Lee, C.-Y., Song, S.-R., Russo, A., Artegiani, A., 1996. Adriatic Sea hydrography. Sci. Mar. 60 (2), 33–43.
Kao, S.-J., 2006. Sapropel burn-down and ichnological response to late Quaternary Sanvoisin, R., D'Onofrio, S., Lucchi, R., Violanti, D., Castradori, D., 1993. 1 Ma
sapropel formation in two ∼400 ky records from the eastern Mediterranean Sea. palaeoclimatic record from the Eastern Mediterranean-Marflux project: first results
Palaeogeogr. Palaeoclim., Palaeoecol. 239, 406–425. of a micropaleontological and sedimentological investigation of a long piston core
Marsaglia, K.M., Fukusawa, H., Cornell, W.C., Skilbeck, C.G., Meyers, P.A., Prasad, M., Klaus, A., from the Calabrian Ridge. Il Quaternario 6, 169–188.
2004. Eustatic Signals in Deep-Marine Sedimentary Sequences Recovered at ODP Site Schmiedl, G., Hemleben, C., Keller, J., Segl, M., 1998. Impact of climatic changes on the
978,AlboranBasin,WesternMediterraneanSea.J.ofSedim.Res.74,378–390. benthic foraminiferal fauna in the Ionian Sea during the last 330,000 years.
Martinez-Ruiz, F., Paytan, A., Kastner, M., González-Donoso, J.M., Linares, D., Bernasconi, Paleoceanography 13 (5), 447–458. doi:10.1029/98PA01864.
S.M., Jimenez-Espejo, F.J., 2003. A comparative study of the geochemical and Sigl, W., Chamley, H., Fabricius, F., Giroud d'Argoud, G., Müller, J., 1978. Sedimentology
mineralogical characteristics of the S1 sapropel in the western and eastern and environmental conditions of sapropels. Initial Reports of the Deep-Sea Drilling
Mediterranean. Palaeogeogr. Palaeoclim. Palaeoecol. 190, 23–37. Project, 42A, pp. 445–465.
Martinson, D.G., Pisias, N.G., Hays, J.D., Imbrie, J., Moore, T.C., Shackleton, N.J., 1987. Age Stefani, C., 2002. Variation in terrigenous supplies in the Upper Pliocene to recent
dating and the orbital theory of the ice ages: development of a high-resolution 0– deposits in the Venice area. Sediment. Geol. 153, 43–55.
300,000-year chronostratigraphy. Quat. Res. 27, 1–29. Stow, D.A.V., Huc, A.-Y., Bertrand, P., 2001. Depositional processes of black shales in deep
Mitchum Jr., R.M., Van Wagoner, J.C., 1991. High-frequency sequences and their stacking water. Mar. Pet. Geol. 18, 491–498.
patterns: sequence stratigraphic evidence of high-frequency cycles. Sedim. Geol. Thunell, R.C., Williams, D.F., 1989. Glacial–Holocene salinity changes in the Mediterra-
70, 131–160. nean Sea: hydrographic and depositional effects. Nature 338, 493–496.
Murat, A., 1999. Pliocene–Pleistocene occurrence of sapropels in the Western Trincardi, F., Cattaneo, A., Asioli, A., Correggiari, A., Langone, L., 1996. Stratigraphy of the
Mediterranean Sea and their relation to eastern Mediterranean sapropels. In: late-Quaternary deposits in the central Adriatic basin and the record of short-term
Zahn, R., Comas, M.C., Klaus, A. (Eds.), Proc. ODP, Sci. Results, 161: College Station, TX climatic events. In: Oldfield, F., Guilizzoni, P. (Eds.), Palaeoenvironmental Analysis
(Ocean Drilling Program), pp. 519–527. doi:10.2973/odp.proc.sr.161.244.1999. of Italian Crater Lake and AdriaticSediments, Mem. Istit. It. Idrobiol., 55, pp. 39–70.
Murat, A., Got, H., 2000. Organic carbon variations of the eastern Mediterranean Vail, P.R., Mitchum Jr., R.M., Thompson III, S., Todd, R.G., Sangree, J.B., Widmier, J.M.,
Holocene sapropel: a key for understanding formation processes. Palaeogeogr. Bubb, J.N., Hatelid, W.G., 1977. Seismic Stratigraphy and Global Sea Level Changes,
Palaeoclimatol. Palaeoecol. 158, 241–257. 11-part contribution to Stratigraphic Interpretation of Seismic Data. In: Payton, C.E.
Negri, A., Capotondi, L., Keller, J., 1999. Calcareous nannofossils and planktic (Ed.), AAPGMemoir, 26, pp. 49–212.
foraminifers distribution patterns in Late Quaternary-Holocene sapropel of the Vail, P.R., Hardenbol, J., Todd, R.G., 1984. Jurassic unconformities, chronostratigraphy
Ionian Sea. Mar. Geol. 157 (1-2), 89–103. and sea-level changes from seismic stratigraphy and biostratigraphy. In: Schlee, J.S.
Olausson, E., 1961. Studies of deep-sea cores Rep. Swed. Deep Sea Exped. 8, 353–391 (Ed.), Interregional Unconformities and Hydrocarbon Accumulation, 36AAPG
1947-1948. Memoir, pp. 129–144.
Passier, H.F., Bosch, H.-J., Nijenhuis, I.A., Lourens, L.J., Böttcher, M.E., Leenders, A., Van Wagoner, J.C., Posamentier, H.W., Mitchum, R.M., Vail, P.R., Sarg, J.F., Loutit, T.S.,
Sinninghe Damsté, J.S., de Lange, G.J., de Leeuw, J.W., 1999. Sulphidic Mediterranean Hardenbol, J., 1988. An overview of sequence stratigraphy and key definitions. In:
surface waters during Pliocene sapropel formation. Nature 397, 146–149. Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamentier, H.W., Ross, C.A., Van
Paterne, M., Guichard, F., Laberyrie, J., Gillot, P.Y., Duplessy, J.C., 1986. Tyrrhenian Sea Wagoner, J.C. (Eds.), Sea Level Changes––An Integrated Approach, 42. SEPM Spec.
Tephrochronology of the Oxigen Isotope Record for the Past 60000 Years. Mar. Geol. Publ. , pp. 39–45.
72, 259–285. Van Wagoner, J.C., Mitchum, R.M., Campion, K.M., Rahmanian, V.D., 1990. Siliciclastic
Paterne, M., Guichard, F., Laberyrie, J., 1988. Explosive activity of the South Italian sequence stratigraphy in well logs, cores, and outcrops. AAPG Methods in
volcanoes during the past 80000 years as determined by marine tephrochronology. Exploration 7, Tulsa, 55.
J. Volcanol. Geotherm. Res. 34, 153–172. Verhallen, P.J.J.M., 1987. Early development of Bulimina marginata in relation to paleo-
Pedersen, T.F., Calvert, S.E., 1990. Anoxia vs. productivity: what controls the formation of environmental changes in the Mediterranean Pliocene. K. Ned. Akad. Wet., Ser. B 90,
organic-carbon-rich sediments and sedimentary rocks? AAPG Bull. 74, 454–466. 161–180.
Piva, A., Asioli, A., Schneider, R.R., Trincardi, F., Andersen, N., Colmenero-Hidalgo, E., Vergnaud-Grazzini, C., Ryan, W.B.F., Cita, M.B., 1977. Stable Isotopic fractionation,
Dennielou, B., Flores, J.-A., Vigliotti, L., 2008a. Climatic cycles as expressed in climate change and episodic stagnation in the eastern Mediterranean during the
sediments of the PROMESS1 borehole PRAD1-2, central Adriatic, for the last 370 ka: Late Quaternary. Marine Micropal. 2, 353–370.
1. Integrated stratigraphy. Geochem. Geophys. Geosyst. 9, Q01R01. doi:10.1029/ von Grafenstein, R., Zahn, R., Tiedemann, R., Murat, A., 1999. Planktonic δ18O records at
2007GC001713. Sites 976 and 977, Alboran Sea: stratigraphy, forcing, and paleoceanographic
Piva, A., Asioli, A., Andersen, N., Grimalt, J.O., Schneider, R.R., Trincardi, F., 2008b. implications. In: Zahn, R., Comas, M.C., Klaus, A. (Eds.), Proc. ODP, Sci. Results, 161:
Climatic cycles as expressed in sediments of the PROMESS1 borehole PRAD1-2, College Station, TX (Ocean Drilling Program), pp. 469–479. doi:10.2973/odp.proc.
central Adriatic, for the last 370 ka: 2. Paleoenvironmental evolution. Geochem. sr.161.233.1999.
Geophys. Geosyst. 9, Q03R02. doi:10.1029/2007GC001785. Waelbroeck, C., Labeyrie, L., Michel, E., Duplessy, J.C., McManus, J.F., Lambeck, K., Balbon,
Posamentier, H.W., Allen, G.P., 1999. Siliciclastic sequence stratigraphy: concepts and E., Labracherie, M., 2002. Sea-level and deep water temperature changes derived
applications. SEPM Concepts in Sedimentology and Paleontology 7, 210. from benthic foraminifera isotopic records. Quat. Sci. Rev. 21, 295–305.
Posamentier, H.W., Jervey, M.T., Vail, P.R., 1988. Eustatic controls on clastic deposition. I. Warning, B., Brumsack, H.-J., 2000. Trace metal signatures of eastern Mediterranean
Conceptual framework. In: Wilgus, C.K., Hastings, B.S., Kendall, C.G.St.C., Posamen- sapropels. Palaeogeogr. Palaeoclimatol. Palaeoecol. 158 (3–4), 293–309.
tier, H.W., Ross, C.A., Van Wagoner, J.C. (Eds.), Sea Level Changes––An Integrated Wignall, P.B., 1994. Black shales. Geology and Geophysics Monographs, vol. 30. Oxford
Approach, SEPM Spec. Publ., 42, pp. 110–124. University Press, Oxford, p. 130.
Ridente, D., Trincardi, F., 2002. Eustatic and tectonic control on deposition and lateral Zahn, R., Sarnthein, M., Erlenkeuser, H., 1987. Benthic isotope evidence for changes of
variability of Quaternary regressive sequences in the Adriatic basin. Mar. Geol. 184, the Mediterranean outflow during the late Quaternary. Paleoceanography 2,
273–293. 543–559.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Benthic environmental changes in the Eastern Mediterranean Sea during


sapropel S5 deposition
Caterina Morigi ⁎
Dipartimento di Scienze del Mare, Università Politecnica delle Marche, Via Brecce Bianche, 60131 Ancona, Italy

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

Article history: High resolution benthic foraminiferal analyses of sediment cores collected at different sites in the Eastern
Received 29 November 2007 Mediterranean Sea revealed that profound changes occurred in deep-sea ventilation over the past 124–119 ka
Received in revised form 16 October 2008 BP. The three cores (SIN97-GC01, BAN89-GC09, BD02-GC01) were selected in order to investigate middle and
Accepted 21 October 2008
deep bathyal ecosystems in the Eastern Mediterranean basin. At site SIN97-GC01 (933 m water depth, Urania
Basin), the presence of benthic fauna in the sapropel layer, and the gradual increase in benthic foraminiferal
Keywords:
Benthic foraminifera
abundance from the middle part of the sapropel, suggest that the sea-floor was partially ventilated and that
Oxygen concentration re-oxygenation had increased considerably during the late phase of S5 deposition. Cores BAN89-GC09
Sapropel (2011 m water depth, Napoli Dome) and BD02-GC01 (2470 m water depth, Ionian Sea) show a different
Mediterranean Sea pattern. During sapropel S5, benthic abundance and diversity strongly decrease and microfauna even
Eemian disappears in some levels, suggesting the establishment and maintenance of stagnant and anoxic conditions
at the seafloor until the end of S5 deposition. In the deepest part of the basin, the re-establishment of the
benthic foraminiferal community at the top of the S5 layer indicates a relatively slow bottom re-oxygenation.
At the southernmost site, the foraminiferal assemblage records a short oxygenation pulse during the
deposition of sapropel S5 linked to a short cold spell. The results of this study suggest that the evolution of
the dysoxic–anoxic conditions, as well as the re-oxygenation pattern at the end of the stagnant period, were
characterised by spatial and temporal variability, possibly controlled by basin physiography.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Although benthic foraminiferal records yield important information,


few papers deal with the assemblages of sapropel S5 (Parker, 1958;
Sapropels are dark organic-rich layers that were deposited Oggioni and Zandini, 1987; Nolet and Corliss, 1990; Vismara-Schilling
cyclically during the last 10 Myr in the Mediterranean Sea. Increasing and Coulbourn, 1991; Schmiedl et al., 2003), and only those of Oggioni
primary production, stratification of superficial waters, and sluggish and Zandini (1987) and Vismara-Schilling and Coulbourn (1991)
bottom-water ventilation are the mechanisms leading to the forma- describe the finer fraction of the sediment (N63 µm). The benthic
tion of these layers. Although several hypotheses have been proposed, faunal succession across sapropel S5 is summarised and re-inter-
it is not yet well understood which of these mechanisms was most preted by Jorissen (1999) who suggests that one of the factors
important, or the precise way in which they led to anoxia in the controlling benthic assemblages is the time between the onset of the
bottom waters. One of the best developed sapropel, termed S5, was oxygen depletion and the re-oxygenation of the bottom environment.
deposited during the last interglacial period (124–119 ka BP, see The paleoceanography of the Eastern Mediterranean Sea during
stratigraphic framework paragraph for more details). Sapropel S5 the last interglacial cycle has been reconstructed from proxies such as
formed during the insolation monsoon maximum of the last planktonic microfossils, organic biomarkers and oxygen and carbon
interglacial period and was interrupted by a dry interlude of several isotopes (Cane et al., 2002; Rohling et al., 2002, 2006; Capotondi et al.,
centuries, with corresponding cooling over the north of the 2006, Giunta et al., 2006; Sangiorgi et al., 2006; Marino et al., 2007). In
Mediterranean area (Rohling et al., 2002). This sapropel was particular, water column dynamics were reconstructed using proxies
characterised by a high organic carbon and biogenic silica accumula- linked to the pelagic environment (Rohling et al., 2006; Marino et al.,
tion rate and by the presence of laminations characterised by 2007).
diatomaceous laminae intercalated with thick layers of mineralogenic Results derived from planktonic microfossils (foraminifera, calcar-
detritus (Kemp et al., 1999; Corselli et al., 2002; Giunta et al., 2006). eous nannofossils, diatoms and organic-walled dinoflagellate cysts)
and geochemistry of the same cores have been previously published
⁎ Present address: Geological Survey of Denmark and Greenland (GEUS), Department
by Capotondi et al. (2006), Giunta et al. (2006) and Sangiorgi et al.
of Stratigraphy, Øster Voldgade 10, 1350 Copenhagen, Denmark. Tel.: +45 3814 2727. (2006) with the aims, respectively, to reconstruct a detailed strati-
E-mail addresses: c.morigi@univpm.it, cmor@geus.dk. graphic framework and to characterise the pelagic paleoecosystem

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.10.010
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 259

dynamics and the paleoceanographic evolution of the superficial water depth) associated with nine other calcareous species (Szarek
water masses during the sapropel S5. et al., 2007). Gooday and Lambshead (1989) noted that Parafissurina
The present study focuses on a high resolution analysis of the fusiformis typically inhabits phytodetritus aggregates in the Porcupine
distribution of benthic foraminifera across sapropel S5 from three Seabight (1345 m water depth). This species is also found in
Eastern Mediterranean Sea cores. The main aim is to describe the association with G. subglobosa, E. exigua and A. weddellensis at abyssal
impact of short-term climatic and environmental changes on different depths south of New Zealand in an area bathed by a cold, corrosive,
deep-sea ecosystems through the study of proxies related to the oligotrophic water mass (Hayward et al., 2007). Furthermore, Fissur-
benthic environment. With this goal in mind, three sites were selected ina spp. constitutes an important part of the recent benthic
to investigate spatial and temporal variability during sapropel microfauna on the hypohaline Black Sea shelf (Yanko, 1990).
deposition as recorded by the benthic microfauna. Globocassidulina subglobosa is another widely distributed species
associated with phytodetritus (Gooday, 1993). Suhr et al. (2003)
1.1. Ecological significance of the benthic foraminifera microfauna reported that it plays an important role in the rapid cycling of
phytoplankton-derived organic carbon on the Antarctic Peninsula
Few studies deal with the ecology of living benthic foraminifera in shelf. In the Atlantic Ocean, G. subglobosa is most abundant in sandy
the deepest part of the Eastern Mediterranean Sea (Parker, 1958; Cita sediments and in areas with relatively strong bottom currents
and Zocchi, 1978; Murray, 1991; De Rijk et al., 1999; De Rijk et al., (Mackensen et al., 1993, 1995; Schmiedl et al., 1997) and low
2000; Pancotti et al., 2008). On the other hand, much has been learnt seasonality in surface primary production (Sun et al., 2006). Hayward
in recent years about the ecology of deep-sea benthic foraminifera in et al. (2007) found this species associated with a low flux of organic
general and our knowledge of the ecological requirements of matter to the seafloor in an oxic environment dominated by strong
important taxa has improved considerably (among others Gooday, currents.
1988; Gooday and Rathburn, 1999; Schmiedl et al., 2000; Fontanier Gyroidinoides spp. is a deep-water taxon, generally associated with
et al., 2002, 2005; Shepherd et al., 2007; see also Gooday, 2003 and low-diversity assemblages related to low organic flux in recent
Jorissen et al., 2007 for general reviews of foraminiferal ecology). samples from the Mediterranean Sea (De Rijk et al., 1999).
Furthermore, studies of the evolution of benthic foraminiferal A first group, lumped as miliolids, comprises large, thick-walled,
assemblages during sapropel deposition have increased our under- epifaunal genera (Pyrgo spp., Spiroloculina spp., Triloculina spp. and
standing of how these organisms respond to oxygen depletion Quinqueloculina spp.). These taxa inhabit well-oxygenated lower
(Jorissen, 1999; Schmiedl et al., 2003). Information about the bathyal and abyssal area of the modern Eastern Mediterranean Sea
ecological requirements of the species found in the present study is and are adapted to a range of oligotrophic to mesotrophic environ-
summarised below. Genera and species are reported in alphabetical ments (Kaiho, 1999; De Rijk et al., 1999; Geslin et al., 2004). Several
order, with exception of selected groups. papers dealing with sapropel deposits indicate that this group is
Articulina tubulosa is not included with other miliolids (see below) characteristic of oligotrophic and well-oxygenated environments and
because it has a slightly different response to increased organic flux. In rapidly disappears when oxygen concentrations fall below critical
modern sediments, it is considered to be an oligotrophic taxon (De values (Jorissen, 1999; Kuhnt et al., 2007).
Rijk et al., 1999), limited to labile organic carbon flux regimes of b2.5 g A second group, the phytodetritus taxa, consists of small rotaliids
C m− 2 y− 1 (De Rijk et al., 2000). In post-sapropel deposits this species that tend to have more or less opportunistic life styles. Normally, they
occurs during the first phase of the re-oxygenation, when the deep- inhabit epifaunal or shallow infaunal microhabitats and occur in
sea environment is still food-enriched (Nolet and Corliss, 1990; relatively eutrophic (but not hypoxic) regions, although some species
Jorissen, 1999). (e.g. Epistominella exigua) can survive in oligotrophic settings (Gooday,
Bolivina is normally considered a low oxygen tolerant taxon 1988, 1993). These tiny species may reproduce rapidly in response to
(Bernhard, 1992) with a variable microhabitat. Some species are strong pulses of fresh organic matter to the seafloor, as reported by
epifaunal or shallow infaunal (e.g. B. spathulata), whereas other Gooday (1988) in the abyssal North Atlantic Ocean. They appear to
species (e.g. B. dilatata) have a clear infaunal tendency (Barmawidjaja exploit phytodetritus when it accumulates on the seafloor after
et al., 1992; De Stigter et al., 1998). Although different species occupy phytoplankton blooms. Such species include Alabaminella weddellen-
different microhabitats, all authors have linked these species to sis, Anomalinoides minimus and Epistominella exigua (Gooday, 1988,
eutrophic and low oxygen environments. 1993). Included in the same group based on their shell morphology
Cassidulina carinata depends on a relatively high flux of fresh, and occurrence in the same stratigraphic levels are Svratkina
highly nutritive organic matter (De Rijk et al., 2000) and occupies tuberculata and other small rotaliids. However, there is no direct
dysoxic environments on the shelf (Hayward et al., 2007). In the evidence that these species respond to phytodetritus.
upwelling area off Mauritania, it dominates assemblages together The third group comprises Globobulimina spp., Bulimina exilis,
with Bolivina dilatata, Chilostomella oolina and Globobulimina affinis, Cassidulinoides bradyi, Chilostomella oolina, and Fursenkoina spp.
at stations where the organic flux is very elevated (N8 g/m2/year and These taxa normally adopt a deep infaunal microhabitat, can survive
the bottom-water oxygen concentration is less than 3 ml/l (Morigi low oxygen conditions and feed on degraded organic matter at the
et al., 2001). dysoxic–anoxic boundary. Several studies (see Fontanier et al., 2006
Cassidulina crassa is common on the middle slope of the Adriatic for a detailed bibliography) report that Globobulimina is a highly
Sea, and has a shallow infaunal microhabitat (De Stigter et al., 1998). specialised taxon that lives in a stable biogeochemical microhabitat.
This species is one of the dominant taxa in the size fraction 63–150 µm Previous studies have suggested that members of this genus prefer
of recent sediment samples from the 2800-m-deep lower canyon degraded rather than fresh organic material (Fontanier et al., 2002).
station in the Bay of Biscay (Fontanier et al., 2005). Suhr and Pond Nomaki et al. (2005, 2006) noted that Globobulimina affinis responds
(2006) investigated the fatty acid composition of this species collected relatively rapidly to fresh phytodetritus, and is able to ingest fresh
off the Antarctic Peninsula, and reported that it responds rapidly to organic material, but can also consume sedimentary organic material
the deposition of phytoplankton material from surface bloom events. when fresh food is not available. A recent study (Risgaard-Petersen
Very few studies deal with ecology of the genus Fissurina, probably et al., 2006) indicates that Globobulimina pseudospinescens, and
because of its small size and low abundance (e.g. 0.4% of the total possibly other infaunal taxa, are able to respire nitrate, allowing
assemblage in the Bay of Biscay, Fontanier et al., 2005), as well as its them to live in anoxic sediments for prolonged periods. Presumably,
presence over a wide depth and geographical range. Living Parafis- the distribution of deep-infaunal taxa in anoxic sediments confers an
surina spp. were found in the abyssal zone of the Sulu Sea (4400 m advantage in terms of food acquisition, predator avoidance, and/or
260 C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

Fig. 1. Bathymetric map of the Eastern Mediterranean Sea. Locations of the investigated cores are represented by dots. Also given are the locations of cores KS205, BAN81-30, BAN81-23,
ODP 971A, LC21, M40-4/67SL, and GeoTüKL83, (indicated by a star) which have yielded key results for comparison (see text and Table 1 for references).

reduced competition. Furthermore, when oxygen concentrations 234.25 cm. The next 4 cm are characterised by a significant colour
decrease, epifaunal and shallow infauna benthic foraminifera dis- change (light grey mud–transitional layer), returning to pale yellow
appear, and the deep infaunal taxa migrate to the water-sediment mud at the top of the section. Core BD02-GC01, collected beneath the
interface (Jorissen et al., 1995). Western Ionian Gyre, where annual PP is about 125 gC m− 2 yr− 1 (Bosc
et al., 2004), includes sapropel S5, which is 9-cm thick and
2. Materials and methods characterised by sharp lower and upper boundaries. Grey sediment
occurs at the top of the sapropel layer (329–327 cm–transitional layer)
The three cores were collected in different sub-basins of the in this core, the last 15 cm of which are characterised by light reddish
Eastern Mediterranean Sea (Fig. 1 and Table 1). The identification of brown mud. The transitional layer present in the three cores could
sapropel S5 in these cores was based on three main lines of evidence correspond to the oxidised part of the sapropel layer, as already
(Fig. 2): 1) biostratigraphic analysis (Giunta et al., 2006) that allows suggested for core BAN89-GC09 by Sangiorgi et al. (2006).
identification of the MNN21a Emiliana huxleyi zone (Rio et al., 1990); Benthic foraminifera assemblages were investigated in detail and
2) when possible, the presence of the younger sapropels (S4 to S1) and qualitative and quantitative analyses were performed on the N63 µm
3) the dark colour of the sediment. For each core, the lower and upper size fraction. Samples containing abundant microfauna have been split
lithological limits of the sapropel are reported in Table 2. in aliquots, all specimens were counted and the relative abundance of
Core SIN97-GC01 (933 m water depth) was collected in the area of each species and species richness (S) are reported, all counts being
the western Cretan counter-clockwise gyre (Pinardi and Masetti, corrected for splits. Foraminiferal numbers are referred to 1 g of bulk
2000) where annual primary productivity (PP) is 125 gC m− 2 yr− 1 dry sediment (BFN). For the layers where the age-model is well
(Bosc et al., 2004). The sapropel interval exhibits sharp lower and constrained (see paragraph on stratigraphic framework) the Benthic
upper boundaries; 2 cm of grey sediment occurs up to the sapropel Foraminiferal Accumulation Rate (BFAR) is calculated, based on a
(transitional layer), passing to oxidised sediment in the upper part of sediment density of 1.43 g/cm3 for the sapropel layer and 1.58 g/cm3
the studied section. Core BAN89-GC09 was collected south of Crete, for the homogeneous sediment characteristic of oxygenated bottom
under the path of the surface eastward current lapping North Africa, waters (Capozzi, personal communication). Ostracods were also
where modern annual PP is estimated at ~ 105 gC m− 2 yr− 1 (Bosc et al., counted, but not differentiated into species: their abundance is
2004). In this core the 20-cm thick sapropel exhibits sharp lower and expressed as number of specimens per dry gram sediment (ON).
upper boundaries. The sapropel is laminated from 241 cm to Diversity indexes were calculated using PAST software (v. 1.33b)

Table 1
Locality and depth of the cores analysed and used for comparison in the present study

Cores Latitude Longitude Water depth (m) Region Reference


SIN97-GC01 35°49.01'N 22°42.04'E 933 Urania basin area This study
BAN89-GC09 33°37.98'N 24°43.29'E 2011 Napoli Dome This study
BD02-GC01 37°33.90'N 17°48.32'E 2470 Ionian basin This study
KS205 38°11.86'N 18°08.04'E 2384 Ionian basin Cane et al. (2002)
BAN81-23 35°52.03'N 20°27.46'E 3026 Ares Crater area Oggioni and Zandini (1987)
30/2 35°50.05'N 20°50.7'E 2885 Poseidon Plateau Vismara-Schilling and Coulbourn (1991)
ODP971A 33°43'N 24°41'E 2026 South Crete Rohling et al. (2006)
LC21 35°66'N 26°58'E 1522 North-eastern Crete Casford et al. (2003)
M40-4/67SL 34°48.8'N 27°17.8'E 2158 North-western Levantine basin Schmiedl et al. (2003)
ODP967C 34°04.27'N 32°43.53' 2554 Eratosthenes seamount Cane et al. (2002)
KL83 32°36.9'N 34°08.9'E 1433 South-eastern Levantine basin Schmiedl et al. (2003)
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 261

(Cane et al., 2002) the results derived from the three cores presented
in this study have been compared with the 971A-equivalent depth
scale (Capotondi et al., 2006). The relative positions between the
correlation markers proposed by Cane et al. (2002), and the change in
colour from light brown to black at the base of sapropel layers in the
three studied cores, appear to be synchronous. Our chronology for S5
relies on a simple linear interpolation between the age of the onset
and the top of the dark layer, suggesting that within the sapropel S5,
1 cm corresponds to ~ 715 year in SIN97-GC01, ~ 260 year in core
BAN89-GC09 and ~ 555 year in core BD02-GC01. The sedimentation
rate for the post-sapropel interval is based on a linear interpolation
between the top of sapropel S5 and the base of sapropel S4 with an
assigned age of 107 Ka (Bar-Matthews et al., 2000) (see Table 2). It is
impossible to calculate the sedimentation rate before sapropel S5,
because there are not clear events in the three cores: sapropel S6 is not
present and also the isotopic signal necessary to determine the
transition from MIS 6 to MIS 5e is incomplete.

3. Results

Fig. 3 summarises the lithological expression of the sapropel S5 in


the three cores, the number of benthic foraminifera (BFN) and
ostracods (ON) per dry gram of sediment along with the Shannon
index [H'(S)] and Evenness index [E] for the benthic foraminiferal
assemblages. BFAR for each core is also presented. The main species
that characterised the pre-sapropel, sapropel and the post-sapropel
layer are illustrated for each core in Figs. 4 to 6. In each case, single
species are illustrated when these are particularly abundant or when
their ecological significance is useful to determine the evolution of the
sea floor. In other cases, species or genera are grouped based on life
Fig. 2. Schematic lithological log and stratigraphical position of the sapropel S5.
Calcareous nannofossil biostratigraphy is based on Giunta et al. (2006).
strategies, as explained in the Section 1.1.

3.1. Core SIN97-GC01 — Fig. 4 and Table 3


(Hammer and Harper, 2005). The following diversity measures were
The BFN exceeds 2000 specimens in the pre-sapropel interval with
calculated: Shannon–Wiener index (ln) [H(S)], and the Evenness index
exception of some levels, at the bottom of the core and at 296 cm
[E]. Shannon index and Evenness are commonly used measures of
(BFN b 1600). BFN strongly decreases in the sapropel and post-
diversity in ecological and foraminiferal studies. Evenness index [E] is a
sapropel layer (average values of 365 ind. g− 1, SD (standard
measure that quantifies how equally specimens are distributed
deviation) = 131 and 311 ind. g− 1 (SD = 165) respectively), with a
amongst species. Values of the evenness index [E] are constrained
small peak in the transitional layer (793 ind. g− 1). BFAR shows the
between 0 and 1. The less the variation in the abundance of species, the
same trend of the BFN but with an evident increase in the transitional
higher the Evenness index [E]. Benthic Foraminiferal Oxygen Index
layer. More than 128 taxa have been identified, the most common
(BFOI) was calculated following Kaiho (1999) together with the oxygen
species being Globocassidulina subglobosa, Cassidulina crassa, Bolivina
index proposed by Schmiedl et al. (2003).
catanensis, Eponides pusillus and Fissurina spp. In the pre-sapropel
layer, Cassidulina carinata and miliolids dominate the assemblage,
2.1. Stratigraphic framework
Bolivina spp. and phytodetritus taxa are also present. A peak of deep-
infaunal taxa together with decreasing numbers of miliolids char-
Sapropel S5 was deposited during the warm and humid period of
acterise the first centimetres of the sapropel layer, whereas the upper
Marine Isotope SubStage (MIS) 5e, representing fully developed
interglacial conditions characterised by ~ 5 °C rise in Atlantic Ocean
sea surface temperature (Shackleton et al., 2003) and sea level ~6 m
above present (Bard et al., 1996). Slightly different time intervals for Table 2
the deposition of sapropel S5 have been suggested. Bar-Matthews Age versus depth of stratigraphic fixed points. The top and the base of the sapropel S5
et al. (2000), Rohling et al. (2002) and Cane et al. (2002) proposed the indicate the lower and upper lithological limit of the dark layer
interval 124–119 ka BP, Struck et al. (2001) suggested 128–123 ka BP, Event Age Core
and Emeis et al. (2003) 127–122 ka BP. Notwithstanding these (ka BP) SIN97-GC01 BAN89-GC09 BD02-GC01
differences, all authors propose that the deposition of sapropel S5 Base S4 (cm) 107 206 169 304
lasted 5000 years. The base of the sapropel is adjusted to the age given Top transitional phase (cm) 283 230 327
in Table 2, which coincides with maxima in Northern Hemisphere Top S5 (cm) 119 285 234.25 329
summer insolation (Rossignol-Strick and Paterne, 1999). Sapropel Base S5 (cm) 124 292 254 338
Width of the dark layer (cm) 7 19.75 9
deposition is believed to be initiated by a significant hydrographical
S5 Sedimentation rate (cm/ka) 1.4 3.95 1.8
event related to an increasing intensity of moisture due to the Top S5 — Base S4 interval 79 65.25 25
intensification of the summer monsoon, suggesting a very rapid width (cm)
response of the Eastern Mediterranean Sea to enhanced precipitation Top S5 — Base S4 interval 6.58 5.43 2.08
in the surrounding catchments (Emeis et al., 2003). Using the sedimentation rate (cm/ka)
1 cm = years 152 184 480
statistical multiproxy correlation approach for records spanning S5
262
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271
Fig. 3. Photograph of the half cores SIN97-GC01, BAN89-GC09 and BD02-GC01. Note the dark sapropel layer (S5) indicated by the shaded area. The dashed line indicates the upper boundary of the transitional phase. Number of benthic
foraminifera per dry gram sediment (BFN), number of ostracods per dry gram sediment (ON), Benthic Foraminiferal Accumulation Rate (BFAR), Diversity [H'(S)] and Evenness [E] of benthic foraminifera assemblage are shown for each core. The
age intervals provided (at right) are based on linear interpolation between estimated ages of 124 and 119 ka BP for the onset and the end of sapropel S5 based on Rohling et al. (2002).
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 263

Fig. 4. Abundance of selected benthic foraminiferal species in core SIN97-GC01. Phytodetritus and deep-infaunal groups are faunal cluster explained in Section 1.1. Shaded interval
represents the sapropel layer and the dashed line indicates the upper boundary of the transitional phase. The age interval provided (at right) is based on a linear interpolation
between estimated ages of 124 and 119 ka BP for the onset and the end of sapropel S5, based on Rohling et al. (2002).

levels are characterised by increased percentages of Cassidulina crassa, decreases (average of 2.5) during the formation of sapropel S5. At
Fissurina spp., Globocassidulina subglobosa, phytodetritus taxa, and the 284 cm, a small increase of diversity [H'(S) = 3] indicates the end of the
disappearance of infaunal taxa. A peak of Articulina tubulosa is sapropel deposition. In the post-sapropel layer, the average value of
observed in the upper level of the sapropel layer. The transitional the Shannon index is 2.8. Evenness has an opposite trend compared to
layer is characterised by a high percentage of Bolivina spp., the the Shannon index with short-term fluctuation along the whole core
occurrence of deep-infaunal taxa and a peak of Gyroidinoides but with higher values in the sapropel layer (0.76) than in the pre and
laevigatus, together with the phytodetritus taxa. The benthic for- post-sapropel levels (0.71) (Table 3).
aminiferal diversity [H'(S)] varies from 1.8 to 3.1 and exhibits short- In the pre-sapropel layer, the ostracod number (ON) is 243 ind. g− 1
term fluctuations throughout the whole core. Diversity slightly on average, whereas within the sapropel layer ostracods are very rare
264 C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

Fig. 5. Abundance of selected benthic foraminiferal species in core BAN89-GC09. Phytodetritus and deep-infaunal groups are faunal cluster explained in the Section 1.1. Shaded
interval represents the sapropel layer and the dashed line indicates the upper boundary of the transitional phase. The age interval provided (at right) is based on linear interpolation
between estimated ages of 124 and 119 ka BP for the onset and the end of sapropel S5, based on Rohling et al. (2002).

or absent. In the post-sapropel, ON displays a small increase, reaching reaching zero in several samples, but exhibit an important peak in the
an average value of 27 ind. g− 1. transitional layer. After the transitional layer BFAR again decreases.
The dominant species in the pre-sapropel layer are Anomalinoides
3.2. Core BAN89-GC09 — Fig. 5 and Table 4 minimus and Eponides pusillus, with small percentages of Gyroidi-
noides laevigatus, Fissurina spp., and Cassidulina carinata. In the
BFN varies from an average value of 166 ind. g− 1 (SD = 76) in the sapropel layer Bolivina spp. are present, together with G. subglobosa,
pre-sapropel interval to 12 ind. g− 1 (SD = 7) during the sapropel although always in very low abundance. At the top of the sapropel
deposition. In the post-sapropel interval, BFN increases to an average different taxa successively dominate the benthic assemblage. Fissurina
value of 33 ind. g− 1 (SD = 17), with a peak of 67 ind. g− 1 in the spp. first increases together with G. laevigatus, followed by Articulina
transitional layer. BFAR values are low during sapropel deposition, tubulosa and finally Anomalinoides minimus. Diversity [H'(S)] in the
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 265

Fig. 6. Abundance of selected benthic foraminiferal species in core BD02-GC01. Phytodetritus and deep infaunal cluster explained in the paragraph 1.1.Section 1.1. Shaded interval
represents the sapropel layer and the dashed line indicates the upper boundary of the transitional phase. The age interval provided (at right) is based on linear interpolation between
estimated ages of 124 and 119 ka BP for the onset and the end of sapropel S5, based on Rohling et al. (2002).

pre-sapropel layer is on average about 2.6 and strongly decreases in if the total number is very low (average 300 ind. g− 1). They disappear
the sapropel with values b1. The benthic fauna is absent in only one during sapropel deposition but occur again after the end of the dark
sample (244 cm), although the sapropel layer yields only 2–4 species. layer deposition (Table 4).
In the pre-sapropel layer, Evenness is rather low with values around
0.4, but shows a rapid increase in the sapropel layer. Mean post- 3.3. Core BD02-GC01 — Fig. 6 and Table 5
sapropel evenness is higher than in the pre-sapropel layer with [E]
value around 0.8. The BFN varies from an average value of 46 ind. g− 1 (SD = 17) to
The number of ostracods per dry gram sediment in the pre- 2 ind. g− 1 (SD = 2 excluding the maximum value of 22 ind. g− 1) from
sapropel layer has the same trend as the density of foraminifera, even the pre-sapropel to the sapropel layer. In the interval between 335 and
266 C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

330 cm, benthic foraminifera are absent, with exception of cm 331 in which the sapropel is 15 cm thick. BFN values were always very low
where 5 ind. g− 1 are present. At 329 cm, benthic foraminifera reappear (0.8 to 2.4 ind. g− 1) and benthic foraminifera were absent in one
with an average value of 27 ind. g− 1 (SD = 32) in the post-sapropel sample. Species of Bolivina, dominated by B. pseudoplicata, were most
layer. BFAR shows the same trend of BFN. A similar suite of species abundant, together with rotaliids. In both cases, the low resolution of
characterise the pre-sapropel and post-sapropel layers, but they show the sampling means that trends in the density and occurrence of
different abundances: Anomalinoides minimus, Gyroidinoides laeviga- benthic foraminiferal species are not clearly apparent, and a detailed
tus, Quinqueloculina seminulum, and Eponides pusillus are more reconstruction of changes in the sea-floor environment during the
abundant in the pre-sapropel than in the post-sapropel layer. On the transition oxic–dysoxic/anoxic condition is not possible. Nevertheless,
other hand, Globocassidulina subglobosa, Articulina tubulosa and Boli- these papers are important because they document the presence of
vina catanensis are present only in the post sapropel interval. In the benthic foraminifera in the sapropel layer. Moreover, the species found
transitional layer different taxa successively dominate: Fissurina spp. in these previous studies are the same as those identified in the three
first reappears afterward Articulina tubulosa increases together with cores analysed during the present investigation.
Anomalinoides minimus. In pre-sapropel layer benthic foraminiferal Other studies (Parisi, 1983; Nolet and Corliss, 1990; Rohling et al.,
diversity [H'(S)] ranges between 1.9 and 1.1 and is generally lower than 1993; Kuhnt et al., 2007; Abu-Zied et al., 2008) based on coarser size
in core BAN89-GC09. Within the sapropel, diversity drops significantly fraction, reported sapropel levels in which benthic foraminifera were
to a minimum average value of 0.7 in levels that are not barren of absent. These barren levels could be due to the hostile environment or
microfauna and evenness shows an opposite trend (Table 5). they could be artefacts created by the use of a coarser sieve. In core
Ostracods are very rare, with a maximum density of 10 ind. g− 1 in KL83 (1433-m water depth), Schmiedl et al. (2003) reported benthic
the pre-sapropel levels; at 338 cm they disappear and occur again at foraminifera in the first part and the upper half of the S5 sapropel
322 cm. (fraction N125 µm). In the shallowest core of the present study (SIN97-
GC01), the BFN (as well as BFAR) strongly decreases in the sapropel but
4. Discussion the sediments are never devoid of foraminifera; the sapropel-fauna
contains a few specimens of G. subglobosa, C. crassa, Bolivina spp., and
4.1. Comparison with previous studies Fissurina spp. together with subordinated E. pusillus and other small
rotaliids. All the species present in the sapropel layer have an average
Many studies dealing with benthic foraminifera and sapropels size less than 150 µm. In the deepest cores, benthic foraminifera are
have focussed on sieve fractions N125 µm or coarser (among others always very rare and small in size; in core BD02-GC01 the entire S5
Rohling et al., 1997; Kuhnt et al., 2007; Abu-Zied et al., 2008). In interval is barren of benthic foraminifera. However, it can not be
particular, previous studies on foraminiferal assemblages of sapropel excluded that foraminifera occurring in the sapropel layers have been
S5 (see Jorissen, 1999 for a review), were performed on size fractions transported down-slope. Even if it is quite easy to recognise taxa
N150 µm. A recent paper (Schmiedl et al., 2003) described results transported from shallower water, resedimentation events that
based on size fractions N125 µm; the finer size fraction 63–125 µm involve a more limited depth-range are not easy to recognise. As
was analysed for only a few selected levels. In the present study, the suggested by Jorissen (1999), increased faunal densities associated
lower limit of 63 µm is adopted because some small species, such as with sapropels, which are accompanied by an increase in the number
Epistominella exigua, Eponides pusillus, other small rotaliids and of species, should be considered with care as they can result from
bolivinids, could be underestimated if a coarser size fraction is chosen. down-slope transport. In the case of the three cores examined in the
In fact, foraminifera vary enormously in size from minute agglutinated present study, the density and diversity show an abrupt decrease and
spheres (10–30 µm) to giant xenophyophores 10 cm or more (Gooday the specimens found in these layers are well preserved and do not
et al., 1995). The dimensions of the calcareous benthic foraminifera, show evidence of transport. Furthermore, the species present in these
which are the main constituent of fossil assemblages, vary also from particular levels are small rotaliids and Bolivina spp. that normally live
tiny (rotaliids and bolivinids) to meiofaunal dimensions (miliolids and in the deepest part of the basin and are tolerant of low-oxygen
lagenids; Sabbatini et al., 2004). In an environment characterised by concentrations. The composition of the benthic microfauna, the trend
extreme conditions, such as seafloor dysoxia, low salinity or carbonate of BFN and of BFAR, and the lack of obvious resedimentation layers,
corrosive waters, small species can represent an important part of the corroborate the hypothesis that the benthic foraminiferal microfauna
benthic foraminiferal community (Boltovskoy et al., 1991; Shepherd is autochthonous.
et al., 2007). In particular, as suggested by Schröder et al. (1987), the
lower size limit of 63 µm provides a more reliable statistical basis for 4.2. Benthic ecosystem variability in the Eastern Mediterranean Sea
paleoceanographic studies. The use of coarser size fractions involves a pre- during and post-sapropel S5 deposition
significant loss of specimens and tends to create artificial “barren”
zones in sequence dominated by small-sized species (Schröder et al., Several studies have investigated the combined influence of water
1987). Shepherd et al. (2007) suggest that smaller species, concen- depth and food availability on benthic microfaunal composition
trated in finer (63–150 µm) size fractions are important for (Jorissen, 1987; Carney, 1989; De Stigter et al., 1998). De Rijk et al.
assessments of foraminiferal assemblages in dysoxic, organic-rich (1999, 2000) illustrate the relationship between deep-sea foraminif-
environments. eral microfauna and the decreasing food availability with increasing
Oggioni and Zandini (1987) analysed the N63-µm size fraction of water depth in the Mediterranean Sea. Ecological studies (Gooday,
sapropel S5 in a core (BAN81-23) collected in the Ares Crater at 3026-m 1993, 2003; Gooday et al., 1998) demonstrate that the diversity of
water depth (Fig. 1). They sampled at intervals of 10 cm across sapropel benthic marine assemblages is also strongly dependant on the stability
S5, which has a thickness of 36 cm. In the three samples collected in the of the environment and the timing of the organic input. A stable
sapropel layer, BFN was higher than zero (3 to 8 ind. g− 1) and ecosystem allows a well-diversified community, whereas ecosystems
resedimentation was observed only in the middle sample (21 ind. g− 1). subject to strong fluctuations sustain a less diversified benthic fauna
The dominant species in this interval were Anomalinoides minimus, dominated by opportunistic taxa. The equation of Berger and Wefer
Bolivina pseudoplicata, Eponides tumidulus (= Eponides pusillus), Cassi- (1990) predicts that the amount of organic matter reaching the
dulina subglobosa and Cassidulina crassa. Another paper published by seafloor is proportional to the primary productivity but also depends
Vismara-Schilling and Coulbourn (1991) analysed sapropel S5 in core on the square of the depth. This means that the deeper the basin the
30/2 collected on the Mediterranean Ridge, Poseidon Plateau, at 2885- less organic matter reaches the seafloor. According to the relation
m water depth. Three samples were collected and analysed in this core, proposed by Herguera and Berger (1991) for 1 g of organic matter that
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 267

reaches the seafloor one benthic foraminiferal shell (N150 µm) is foraminifera. Although BFAR values are quite different, similar down-
generated. Thus the BFAR can be used as a proxy for organic flux to the core trends in the three cores imply a similar evolution of the benthic
seafloor, as long as the sedimentation rate is constant. environment: elevated values of BFN in the pre-sapropel levels
Several studies indicate that oligotrophic environments support indicate optimal conditions for benthic life, followed by a strong
fewer benthic foraminifera, but when the organic flux to the sea floor decrease (not reaching zero) when organic carbon started to
increases, BFAR values also increase strongly (Loubere, 1994). Of accumulate (corresponding to the onset of the dark level), and low
course, this equation cannot strictly apply to the present study abundance and/or barren levels in the sapropel sensu strictu. Above
because 1) the size fraction analysed was N63 µm rather than the sapropel, where the sediment colour changes from black to grey, a
N150 µm; 2) it is not possible to calculate BFAR in the pre-sapropel peak of BFAR marks a transitional phase in the three cores. This peak is
interval and 3) low oxygen concentrations have a major impact on the characterised by the re-occurrence of Fissurina spp., followed by Ar-
benthic microfauna (Naidu and Malmgren, 1995). Changes in the flux ticulina tubulosa and phytodetritus taxa. This foraminiferal succession
of organic carbon to the sea-bed causes shifts in the faunal that occurs at the deepest sites at the end of the sapropel deposition
composition. Such shifts have been recognised in the studied cores. indicates that the re-oxygenation was fast enough to allow the
The benthic assemblage composition and faunal trends through the phytodetritus taxa to colonise the sea bottom (Jorissen, 1999). After
studied interval were similar in cores BAN89-GC09 and BD02-GC01. the transitional phase, benthic foraminiferal density remains low, and
Core SIN97-GC01, on the other hand, yielded different species miliolids increase, indicating a return to oligotrophic and well-
assemblages and the trends in BFN and BFAR were slightly different oxygenated environment. The low BFAR recorded at the three sites
from those observed in the other cores. Different water depths indicates a slow ecosystem recovery. About 5 kyr after the termination
together with different PP and hence different organic flux rates to the of sapropel deposition, the benthic environment totally recovered, as
seafloor (Jz) may be the main causes of this difference between the testified by the high value of BFAR at the top of the studied interval in
sites. In fact, during the oxygenated interval (pre-sapropel), the BFN in core BD02-GC01. Ostracods, however, recover more slowly at the end
core SIN97-GC01 was more than 10 times higher than in core BAN89- of the sapropel deposition and return only after the transitional phase.
GC09 and 50 times higher than in core BD02-GC01. The position of The similar trends of BFN and BFAR recorded in the three sites suggest
core SIN97-GC01 is optimal for the development of the benthic that the seafloor environment in the Eastern Mediterranean was
microfauna due to the shallow location (recent Jz = 234 gC/m2/year characterised by low oxygen concentration during sapropel deposi-
calculated with formula of Berger and Wefer, 1990) and proximity to tion, although with different intensities in the different areas, as
the continental shelf, which supplies additional organic matter. At this discussed in the next section.
site, a rich and diverse benthic foraminiferal assemblage develops. It
consists mainly of bolivinids, Cassidulina carinata, Fissurina spp., 4.3. Evolution of the anoxia at the sea bottom and short term variability
miliolids, phytodetritus taxa, and even includes subordinate numbers
of deep-infaunal taxa. According to the TROX model of Jorissen et al. The isotopic signal of shallow-dwelling Globigerinoides ruber and
(1995), the combination of a moderate flux of organic matter and the absence of deep-dwelling planktonic foraminifera such as Glo-
elevated concentrations of oxygen above and in the sediment results borotalia inflata, indicate that the onset of sapropel S5 deposition
in the occupation of both epifaunal and deep infaunal niches. In the coincided with a rapid increase in surface-water stratification that
deepest cores (BAN89-GC09 and BD02-GC01), foraminiferal densities inhibited the deep-water formation and resulted in anoxic condition
in the pre-sapropel sediments are lower with a high percentage of G. in the deep Mediterranean basins (Rohling et al., 2006). Although,
laevigatus and phytodetritus taxa that indicate an oligotrophic, well- benthic foraminifera, as a group, are extremely tolerant of severe
oxygenated environment. The positive correlation between benthic oxygen depletion and can survive long periods of dysoxia/anoxia
foraminiferal density and organic flux to the seafloor applies until the (Josefson and Widbom, 1988; Bernhard and Reimers, 1992; Risgaard-
oxygen level drops below a critical value (Naidu and Malmgren, 1995; Petersen et al., 2006), they were strongly impacted by this long-term
Morigi et al., 2001). Only specialised taxa can survive in dysoxic oxygenation crisis. Across the investigated intervals, BFAR (and BFN)
conditions, and hence, if dysoxia or anoxia persist, BFAR dramatically values exhibit strong fluctuations that can be attributed to extensive
decreases and benthic foraminifera eventually disappear. A drop in the and drastic decreases in oxygen concentrations at the seafloor (Fig. 2).
abundance of miliolids before the onset of sapropel deposition, In particular, in the deepest core (BD02-GC01), the absence of benthic
combined with a decreases in BFAR and BFN and an increase of low- microfossils indicates persistent anoxia during the time of the
oxygen indicators (infaunal taxa), suggest a strong decrease in oxygen sapropel deposition. On the other hand, some intervals in core
concentrations. The peak of species, such as Globobulimina spp. and BAN89-GC09 show a very low BFAR (and BFN) value and in the
Chilostomella spp., which consume degraded rather than fresh organic shallowest core SIN97-GC01 the sediment is never barren of benthic
material, suggests increasing preservation of the organic matter at the foraminifera. To illustrate the changes in oxygen concentrations
seafloor, favouring the development of faunas feeding on low-quality during the deposition of the sapropel S5, the Benthic Foraminiferal
food sources. In both the deepest cores, there is a strong increase in Oxygen Index (BFOI) of Kaiho (1999), and the oxygen index proposed
the abundance of phytodetritus taxa after the peak of deep-infaunal by Schmiedl et al. (2003), were calculated for the three cores (Fig. 7).
taxa and before the disappearance of the benthic microfauna. The These paleo-oxygenation proxies are discussed in detail by Jorissen
presence of phytodetritus taxa suggests that pulsed inputs of labile et al. (2007), who highlight some pitfalls for both indexes. One of the
organic matter triggered seafloor oxygen depletion. main problems is that all the species respond to organic flux increases
A previous study (Weldeab et al., 2003) estimated a Jz value of as well as decreases in bottom-water oxygenation. Furthermore
350 gC m− 2 yr− 1 during the deposition of sapropel S5 in the area of Kahio's index is based on oxic, suboxic and dysoxic indicators that
core BAN89-GC09, about three times higher than present values are not calibrated with respect to recent faunas. It is therefore
(PP = 106 gC m− 2 yr− 1, Jz = 123 gC m− 2 yr− 1). This corresponds to a very important to remember that, in the three cores analysed during the
high export production, similar to that reported in modern upwelling present study, variations in these two oxygen indexes are always
areas (Berger, 1989). At the shallower site (SIN97-GC01), the presence accompanied by a strong decrease in the number of foraminiferal tests
of phytodetritus species again points to pulsed organic-matter inputs, per gram dry sediment, and that the sapropel levels contain a low-
while the low abundance of taxa feeding on degraded material diversity fauna, strongly dominated by opportunistic taxa (phytode-
indicates higher oxygen concentrations. The disappearance of ostra- tritus taxa or deep infauna).
cods with decreasing seafloor oxygen concentrations suggests that The two indexes suggest that suboxic condition (oxygen content
these metazoans are more sensitive to oxygen depletion than benthic between 0.3and 1.5 ml/l) developed abruptly. The sudden
268 C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

Fig. 7. Records through sapropel S5 of the benthic foraminiferal oxygen index (BFOI) (Kahio, 1999) and the index proposed by Schmiedl et al. (2003). The grey scale shows the
empirical relation between BFOI and dissolved oxygen levels, as calculated by Kaiho (1999). The stable oxygen isotope record for the planktonic foraminifera Globigerinoides ruber is
provided by Capotondi et al. (2006). Grey blocks indicate the visual extent of the dark coloured sapropel sediments in each core, dashed line indicates the upper boundary of the
transitional phase and the solid lines represent primary correlation levels between the studied sites (Capotondi et al., 2006) and the stratigraphic framework proposed by Cane et al.
(2002). The stratigraphic marker i6 corresponds to the prominent shoulder before sharp enrichment trend of δ18O G. ruber and stratigraphic marker f4 corresponds to the
disappearance of Globorotalia scitula in the planktonic assemblage.

disappearance of oxic species, combined with the strong decrease of indexes in core BAN89-GC09 yield inconsistent results in the sapropel
BFAR (BFN) values, reflects the development of the dysoxic/anoxic layer: Schmiedl's index does not show strong fluctuations during
conditions in the Mediterranean Sea. Comparison between the sapropel deposition whereas BFOI shows a more variable trend. In fact,
records from the deepest cores suggests that anoxia developed first in several levels the presence of oxic/dysoxic indicators together with
in the southern Mediterranean area (core BAN89-GC09) and then in a small peak of BFAR, corresponding to an increase of BFOI, suggests
the Ionian Sea (core BD02-GC01). Considering the sedimentation rate that the input of oxygenated water at the seafloor led to the re-
calculated for the two cores (see Table 2), there is a gap of about 500 yr establishment of a benthic population. In core BAN89-GC09, which
between the development of anoxia at the site of BAN89-GC09 and at has the highest sedimentation rate (3.95 cm/ka), a repopulation event
the site of core BD02-GC01. This delay in the onset of persistent anoxia involving the reappearance of oxic indicators suggests that there was a
between the two cores could be due to several factors. First, there is a marked variability in oxygen concentrations. This event, which is
difference in water depth between the cores. One hypothesis is that confirmed by a BFOI peak in the interval 249–246 cm (oxygen content
the delayed onset of anoxia at BD02-GC01 (the deepest site) could N3.2 ml/l, on the base of Kahio index), could indicate instability of the
reflect a slow downward expansion of the oxygen minimum zone, as water column or an abrupt injection of deep oxygenated water.
proposed by the model of Strohle and Krom (1997) and Stratford et al. Considering a depositional period of 5 ka for the duration of S5, this
(2000). A second hypothesis is that anoxia developed later in the short event lasted about 800 years. Although it left no lithological
northernmost sector. Cane et al. (2002) observed a delay in sapropel trace in the sapropel layer, other proxies can be used to identify this
deposition in the easternmost sector (ODP976C — Eratosthenes short episode. In particular, it corresponds to a peak in the abundance
seamount) and suggested that this time difference could be due to of the deep dwelling planktonic foraminifera Globorotalia scitula
lower productivity/export production, or to a period of occasional (Capotondi et al., 2006). In the stratigraphic framework proposed by
deep water ventilation. The site of core BD02-GC01 is affected at Cane et al. (2002), the disappearance of this species is indicated as
present by injection of oxygenated deep-water formed in the Adriatic correlation point f4 (Fig. 7). Rohling et al. (2002) reported that the
Sea. This process could have influenced this site also during the time of peak coincides with a δ 18OG.ruber anomaly in the Ionian Sea and in the
sapropel S5 deposition and delayed the onset of anoxia. When south Mediterranean Sea. According to these authors, this depletion in
bottom-water ventilation collapsed in the Ionian Sea, anoxia persisted the δ18OG.ruber indicates a relaxation of the northward penetration of
until 124 ka BP, whereas in the south Mediterranean Sea (core BAN89- the Intertropical Convergence Zone (ITCZ) that limits the duration of
GC09) the situation may have been more variable. The two oxygen the monsoon to 1–2 month in the area of the Libyan basin. The oxygen
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 269

isotope depletion is also recorded in our cores. Following Cane et al. in response to the change in hydrographic forcing and organic flux to
(2002) the correlation point i6 indicates the prominent shoulder the seafloor. As a result, the seafloor remained anoxic for the entire
before the sharp enrichment trend of δ18OG.ruber(Fig. 7). In our record, duration of the sapropel deposition. However, a short, cold event
this oxygen isotope depletion corresponds to the increase in BFOI, recorded in the southern part of the Mediterranean Sea indicate
although this event is not reflected by the Schmiedl Index. Sangiorgi instability in the water column and a partial re-oxygenation of the
et al. (2006) found evidence for it in the same core using a multiproxy seafloor that allowed brief colonisation of benthic foraminiferal
approach and suggest that it reflects upper water column mixing microfauna. At the end of the sapropel deposition, during the
together with a decrease in temperature and an increase in seasonal transitional phase, the oxygen concentration increased strongly as
contrast. This event is not evident in core BD02-GC01. At this site, indicated by the presence of taxa that require a high oxygen
anoxic conditions that persisted until the end of sapropel deposition concentration. The environment did not return to its pre-sapropel
and/or the low sedimentation rate (1.8 cm/ka) could mask the record condition, or improve enough to allow a total re-colonisation of the
of this brief event. seafloor.
A different situation is represented in core SIN97-GC01, where a Analyses of the 63 µm fractions yielded more information on the
low BFAR in the sapropel layer is represented by a mixture of oxic and benthic assemblage during stress episodes than coarser fractions,
dysoxic species. Both the BFOI and the Schmiedl index indicate that making it possible to evaluate benthic ecosystem variability. In fact,
the bottom water never became anoxic and the presence of benthic small taxa dominated foraminiferal assemblages when the environ-
fauna demonstrates that oxygen was consistently available at 933 m mental conditions at the seafloor were unfavourable to the benthic
depth, although at low concentration. The occurrence of benthic life. In particular, Fissurina, a genus that normally is not considered in
foraminifera in the late phase of sapropel S5 formation was already this type of study due to its small size and poorly-known ecological
described by Schmiedl et al. (2003) in a core collected near the Israeli requirements, seems to play an important role in the re-colonisation
margin at a water depth of 1433 m. These authors suggest that a of the seafloor at the end of the sapropel deposition and is the first
gradual cooling may have favoured the formation of dense water at taxon to reappear during the transitional phase after the sapropel
intermediate and deep formation sites. The continuous presence of deposition.
benthic foraminifera in core SIN97-GC01 through sapropel S5 implies
that persistent anoxia did not develop in that region. In the modern
Acknowledgements
Black Sea, which is anoxic below 150 m depth, several authors
(Codispoti et al., 1991; Özsoy and Ünlüata, 1997) have described the
Special thanks go to S. Giunta, A. Negri and A. Sabbatini for their
structure of the chemocline as a dome, i.e. shallower in the middle of
constructive comments and discussions. I would like to thank warmly,
the basin and deeper near the shelf. Rohling et al. (2006) estimate the
A. Gooday which significantly helped to improve the manuscript. E.
depth range of the chemocline in the middle of the Eastern
Thomas and an anonymous reviewer are kindly acknowledged for
Mediterranean Sea to have been 150–300 m during S5 deposition
their helpful and critical review of the manuscript. This research is
and suggest that the cyclonic circulation may have caused the
part of the project “The dynamic of the Mediterranean paleoecosys-
chemocline to reside considerably deeper near the margins of the
tem during the Eemian” (Coord. C. Corselli) funded by the Italian
basin. A chemocline at about 1000 m water depth is needed to explain
MIUR.
the presence of a stressed benthic microfauna in the shallower core. A
second hypothesis is supported by the model of Maldonado and
Stanley (1976), who suggested that a partly ventilated water mass Appendix A. Supplementary data
existed between 1000 and 2000 m depth and that the deepest part of
the basin was occupied by stagnant and anoxic water. Finally, a third Supplementary data associated with this article can be found, in
hypothesis was proposed by Casford et al. (2003), who suggested that the online version, at doi:10.1016/j.palaeo.2008.10.010.
intermittent bottom-water ventilation occurred throughout periods of
sapropel deposition, creating high-frequency variability in deep-water References
formation. These authors proposed that a blanket of anoxia developed
under areas of higher productivity, and frequent intermittent bottom- Abu-Zied, R.H., Rohling, E.J., Jorissen, F.J., Fontanier, C., Casford, J.S.L., Cooke, S., 2008.
Benthic foraminiferal response to changes in bottom-water oxygenation and
water ventilation occurred throughout periods of sapropel deposition.
organic carbon flux in the eastern Mediterranean during LGM to Recent times. Mar.
In area of low productivity, there was no bottom-water anoxia, but Micropaleontol. 67, 46–68.
poor ventilation created dysoxic conditions, allowing the observed Bard, E., Jouannic, C., Hamelin, B., Pirazzoli, P., Arnold, M., Faure, G., Sumosusastro, P.,
persistence of specialised benthic foraminifera. Syefudin, P.A., 1996. Pleistocene sea levels and tectonic uplift based on radiometric
dating of coral terraces at Sumba Island, Indonesia. Geophys. Res. Lett. 23,
It is impossible to decide which, if any, of these hypotheses is 1473–1476.
correct based on benthic foraminiferal data from only three cores. Barmawidjaja, D.M., Jorissen, F.J., Puskaric, S., Van der Zwaan, G.J., 1992. Microhabitat
However, the data reported here suggest that although the overall selection by benthic Foraminifera in the northern Adriatic Sea. J. Foraminiferal Res.
22 (4), 297–317.
ventilation in the Eastern Mediterranean Sea was much reduced, Bar-Matthews, M., Ayalon, A., Kaufman, A., 2000. Timing and hydrological conditions of
regional differences existed in the intensity and frequency of the Sapropel events in the Eastern Mediterranean, as evident from speleothems, Soreq
bottom-water oxygen depletion. cave, Israel. Chem. Geol. 169, 145–156.
Berger, W.H., 1989. Global maps of ocean productivity. In: Berger, W.H., Smetacek, V.S.,
Wefer, G. (Eds.), Productivity of the Ocean: Present and Past. John Wiley, pp. 429–455.
5. Conclusions Berger, W.H., Wefer, G., 1990. Export production: seasonality and intermittency, and
paleoceanographic implications. Palaeogeogr. Palaeoclimatol. Palaeoecol. 89,
245–254.
The onset of sapropel deposition is normally considered to be Bernhard, J.M., 1992. Benthic foraminiferal distribution and biomass related to pore-
coincident with the disappearance of benthic foraminifera. However water oxygen content: Central California continental slope and rise. Deep-Sea Res.
the studied cores reveal a different situation characterised by high 39, 585–605.
Bernhard, J.M., Reimers, C.E., 1992. Benthic foraminiferal population fluctuations related
internal and spatial variability in the sapropel. The onset of the
to anoxia: Santa Barbara Basin. Biogeochem. 15, 127–149.
sapropel deposition is characterised by a strong decrease of the Boltovskoy, E., Scott, D.B., Medioli, F.S., 1991. Morphological variation of benthic
benthic foraminiferal number due to the low-oxygen conditions at the foraminiferal tests in response to changes in ecological parameters: a review.
seafloor. On the Mediterranean margin, poor ventilation created J. Paleontol. 65, 175–185.
Bosc, E., Bricaud, A., Antoine, D., 2004. Seasonal and interannual variability in algal
dysoxic condition allowing the presence of a stressed benthic biomass and primary production in the Mediterranean Sea, as derived from 4 years
microfauna. At the deeper stations ventilation collapsed completely of SeaWiFS observations. Glob. Biogeochem. Cycle 18. doi:10.1029/2003GB002034.
270 C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271

Cane, T., Rohling, E.J., Kemp, A.E.S., Cooke, S., Pearce, R.B., 2002. High resolution Jorissen, F.J., Fontanier, C., Thomas, E., 2007. Paleoceanographical proxies based on deep-
stratigraphic framework for Mediterranean sapropel S5: Defining temporal sea benthic foraminiferal assemblage characteristics. In: Hillaire-Marcel, C., de Vernal,
relationships between records of Eemian climate variability. Palaeogeogr. Palaeo- A. (Eds.), Proxies in Late Cenozoic paleoceanography. Elsevier, pp. 263–313.
climatol. Palaeoecol. 183, 87–101. Josefson, A.B., Widbom, B., 1988. Differential response of benthic macrofauna and
Capotondi, L., Principato, M.S., Morigi, C., Sangiorgi, F., Maffioli, P., Giunta, S., Negri, A., meiofauna to hypoxia in Gullmar Fjord basin. Mar. Biol. 100, 31–41.
Corselli, C., 2006. The Eastern Mediterranean Sapropel S5 interval: foraminiferal Kaiho, K., 1999. Effect of organic carbon flux and dissolved oxygen on the benthic
distribution and stratigraphical implications. Palaeogeogr. Palaeoclimatol. Palaeoe- foraminiferal oxygen index (BFOI). Mar. Micropaleontol. 31, 67–76.
col. 235, 48–65. Kemp, A.E.S., Pearce, R.B., Koizumi, I., Pike, J., Rance, S.J., 1999. The role of mat-forming
Carney, R.S., 1989. Examining relationship between organic carbon flux and deep sea diatoms in the formation of Mediterranean sapropels. Nature 398, 57–61.
deposit feeding. In: Lopez, G., Taghon, G.T., Levinton, J. (Eds.), Ecology of Marine Kuhnt, T., Schmiedl, G., Ehrmann, W., Hamann, Y., Hemleben, Ch., 2007. Deep-sea
Deposit Feeder. Springer–Verlag, New York, pp. 24–58. ecosystem variability of the Aegean Sea during the past 22 ka as revealed by Benthic
Casford, J.S.L., Rohling, E.J., Abu-Zied, R.H., Fontanier, C., Jorissen, F.J., Leng, M.J., Foraminifera. Mar. Micropaleontol. 64, 141–162.
Schmiedl, G., Thomson, J., 2003. A dynamic concept for eastern Mediterranean Loubere, P., 1994. Quantitative estimation of surface ocean productivity and bottom water
circulation and oxygenation during sapropel formation. Palaeogeogr. Palaeoclima- oxygen concentration using benthic foraminifera. Paleoceanography 9, 723–737.
tol. Palaeoecol. 190, 103–119. Mackensen, A., Fütterer, D.K., Grobe, H., Schmiedl, G., 1993. Benthic foraminiferal
Cita, M.B., Zocchi, M., 1978. Distribution patterns of benthic foraminifera on the floor of assemblages from the eastern South Atlantic Polar Front region between 358 and
the Mediterranean Sea. Oceanol. Acta 1 (4), 445–462. 578S: distribution, ecology and fossilization potential. Mar. Micropaleontol. 22,
Codispoti, L.A., Friederich, G.E., Murray, J.W., Sakamoto, C.M., 1991. Chemical variability 33–69.
in the Black Sea: implications of continuous vertical profiles that penetrated the Mackensen, A., Schmiedl, G., Harloff, J., Giese, M., 1995. Deep-sea Foraminifera in the South
oxic–anoxic interface. Deep-Sea Res. 38, 691–710. Atlantic Ocean: ecology and assemblage generation. Micropaleontology 41 (4),
Corselli, C., Principato, M.S., Maffioli, P., Crudeli, D., 2002. Changes in planktonic 342–358.
assemblages during sapropel S5 deposition: evidence from Urania Basin area, Maldonado, A., Stanley, D.J., 1976. Late Quaternary sedimentation and stratigraphy in
eastern Mediterranean. Paleoceanography 17. doi:10.1029/2000PA000536. the Strait of Sicily. Smithson. Contrib. Earth Sci. 16, 1–73.
De Rijk, S., Troelstra, S.R., Rohling, E.J., 1999. Benthic foraminiferal distribution in the Marino, G., Rohling, E.J., Rijpstra, W.I.C., Sangiorgi, S., Schouten, S., Sinninghe Damsté,
Mediterranean Sea. J. Foraminiferal Res. 29 (2), 93–103. J.S., 2007. Aegean Sea as driver of hydrographic and ecological changes in the
De Rijk, S., Jorissen, F.J., Rohling, E.J., Troelstra, S.R., 2000. Organic flux control on eastern Mediterranean. Geology 35, 675–678.
bathymetric zonation of Mediterranean benthic foraminifera. Mar. Micropaleontol. Morigi, C., Jorissen, E.J., Gervais, A., Guichard, S., Borsetti, A.M., 2001. Benthic
40, 151–166. foraminiferal faunas in surface sediments off NW Africa: relationship with organic
De Stigter, H.C., Jorissen, F.J., Van der Zwaan, G.J., 1998. Bathymetric distribution and flux to the ocean floor. J. Foraminiferal Res. 31 (4), 350–368.
microhabitat partitioning of live (Rose Bengal stained) benthic foraminifera along a Murray, J., 1991. Ecology and Paleoecology of Benthic Foraminifera. Longman Scientific
shelf to deep sea transect in the southern Adriatic Sea. Mar. Micropaleontol. 28, & Technical, London. 398 pp.
40–65. Naidu, P.D., Malmgren, B.A., 1995. Do benthic foraminifer records represent a
Emeis, K.-C., Schulz, H., Struck, U., Rossignol-Strick, M., Erlenkeuser, H., Howell, M.W., productivity index in oxygen minimum zone areas? An evaluation from the
Kroon, D., Mackensen, A., Ishizuka, S., Oba, T., Sakamoto, T., Koizumi, I., 2003. Oman Margin, Arabian Sea. Mar. Micropaleontol. 26, 49–55.
Eastern Mediterranean surface water temperatures and δ18O composition during Nolet, G.J., Corliss, B.H., 1990. Benthic foraminiferal evidence for reduced deep-water
deposition of sapropels in the late Quaternary. Paleoceanography 18 (1). circulation during sapropel deposition in the Eastern Mediterranean. Mar. Geol. 94,
doi:10.1029/2000PA000617. 109–130.
Fontanier, C., Jorissen, F.J., Licari, L., Alexandre, A., Anschutz, P., Carbonel, P., 2002. Live Nomaki, H., Heinz, P., Nakatsuda, T., Shimanaga, M., Kitazato, H., 2005. Species-specific
benthic foraminiferal faunas from the Bay of Biscay: faunal density, composition, ingestion of organic carbon by deep-sea benthic foraminifera and meiobenthos: in
and microhabitats. Deep-Sea Res. I 49, 751–785. situ tracer experiments. Limnol. Oceanogr. 50 (1), 134–146.
Fontanier, C., Jorissen, F.J., Chaillou, G., Anschutz, P., Grémare, A., Griveaud, C., 2005. Live Nomaki, H., Heinz, P., Nakatsuka, T., Shimanaga, M., Ohkouchi, N., Ogawa, N.O., Kogure,
foraminiferal faunas from a 2800 m deep lower canyon station from the Bay of K., Ikemoto, E., Kitazato, H., 2006. Different ingestion patterns of 13C-labeled bacteria
Biscay: Faunal response to focusing of refractory organic matter. Deep-Sea Res. I 52, and algae by deep-sea benthic foraminifera. Mar. Ecol., Prog. Ser. 310, 95–108.
1189–1227. Oggioni, E., Zandini, L., 1987. Response of benthic foraminifera to stagnant episodes — a
Fontanier, C., Jorissen, F.J., Anschutz, P., Chaillou, G., 2006. Seasonal variability of benthic quantitative study of core BAN81-23, Eastern Mediterranean. Mar. Geol. 75, 241–261.
foraminiferal fauna at 1000 m depth in the Bay of Biscay. J. Foraminiferal Res. 36 (1), Özsoy, E., Ünlüata, Ü., 1997. Oceanography of the Black Sea: a review of some recent
61–76. results. Earth-Sci. Rev. 42, 231–272.
Geslin, E., Heinz, P., Jorissen, F.J., Hemleben, Ch., 2004. Migratory responses of deep-sea Pancotti, I., Sabbatini, A., Anibaldi, A., Gooday, A.J., Morigi, C., Negri, A., 2008. Recent
benthic foraminifera to variable oxygen conditions: laboratory investigations. Mar. benthic foraminifera of West Mediterranean Sea and Gulf of Cadiz. Volume of
Micropaleontol. 53, 227–243. Abstract of HERMES Workshop: Taxonomy and ecology of recent Foraminifera.
Giunta, S., Negri, A., Maffioli, P., Sangiorgi, F., Capotondi, L., Morigi, C., Principato, M.S., Polytechnic University of Marche, p. 20. 7 – 10 April 2008.
Corselli, C., 2006. Isotopic stage 5e in the Eastern Mediterranean Sea: effects of Parisi, E., 1983. Distribuzione dei Foraminiferi bentonici in una carota della Dorsale
nutrient and circulation on phytoplankton. Palaeogeogr. Palaeoclimatol. Palaeoecol. Mediterranea (Pleistocene medio e superiore). Riv. Ital. Paleontol. 88, 641–677.
235, 28–47. Parker, F.L., 1958. Eastern Mediterranean foraminifera. Reports of the Swedish Deep Sea
Gooday, A.J., 1988. A response by benthic foraminifera to phytodetritus deposition in the Expedition, vol. III: Sediment cores from the Mediterranean and the Red Sea, n. 4,
deep-sea. Nature 332, 70–73. pp. 219–283.
Gooday, A.J., 1993. Deep-sea benthic foraminiferal species which exploit phytodetritus: Pinardi, N., Masetti, E., 2000. Variability of the large scale general circulation of the
characteristic features and controls on distribution. Mar. Micropaleontol. 22, Mediterranean Sea from observations and modelling: a review. Palaeogeogr.
187–205. Palaeoclimatol. Palaeoecol. 158, 153–173.
Gooday, A.J., 2003. Benthic foraminifera (Protista) as tools in deep-water palaeoceano- Rio, D., Raffi, I., Villa, G., 1990. In: Kastens, K.A., Mascle, J., et al. (Eds.), Pliocene–
graphy: a review of environmental influences on faunal characteristics. Adv. Mar. Pleistocene calcareous nannofossil distribution patterns in the Western Mediterra-
Biol. 46, 1–90. nean. Proc. of ODP, Sci. Results, vol. 107, pp. 513–533.
Gooday, A.J., Lambshead, P.J.D., 1989. Influence of seasonally deposited phytodetritus on Risgaard-Petersen, N., Langezaal, A.M., Ingvardsen, S., Schmid, M.C., Jetten, M.S., Op den
benthic foraminiferal populations in the bathyal northeast Atlantic. Mar. Ecol., Prog. Camp, H.J., Derksen, J.W., Pina-Ochoa, E., Eriksson, S.P., Nielsen, L.P., Revsbech, N.P.,
Ser. 58, 53–67. Cedhagen, T., van der Zwaan, G.J., 2006. Evidence for complete denitrification in a
Gooday, A.J., Rathburn, A.E., 1999. Temporal variability in living deep-sea benthic benthic foraminifer. Nature 443, 93–96.
foraminifera: a review. Earth-Sci. Rev. 46, 187–212. Rohling, E.J., De Stigter, H.C., Vergnaud, - Grazzini, C., Zaalberg, R., 1993. Temporary
Gooday, A.J., Carstens, M., Thiel, H., 1995. Micro- and Nanoforaminifera from Abyssal repopulation by low-oxygen tolerant benthic foraminifera within an upper Pliocene
Northeast Atlantic sediments: a preliminary report. Int. Rev. Hydrobiol. 80, sapropel: evidence for the role of oxygen depletion in the formation of sapropels.
361–383. Mar. Micropaleontol. 22, 207–219.
Gooday, A.J., Bett, B.J., Shires, R., Lambshead, P.J.D., 1998. Deep-sea benthic foraminiferal Rohling, E.J., Jorissen, F.J., De Stigter, H.C., 1997. 200 Year interruption of Holocene
species diversity in the NE Atlantic and Arabian Sea: a synthesis. Deep-sea Res. II, sapropel formation in the Adriatic Sea. J. Micropaleontol. 16, 97–108.
45, 165–201. Rohling, E.J., Cane, T.R., Cooke, S., Sprovieri, M., Bouloubassi, I., Emeis, K.-C., Schiebel,
Hammer, Ø., Harper, D.A.T., 2005. http://folk.uio.no/ohammer/past. R., Kroon, D., Jorissen, F.J., Lorre, A., Kemp, A.E.S., 2002. African monsoon
Hayward, B.W., Grenfell, H.R., Sabaa, A.T., Neil, H.L., 2007. Factors influencing the variability during the previous interglacial maximum. Earth Planet. Sci. Lett.
distribution of Subantarctic deep-sea benthic foraminifera, Campbell and Bounty 202, 61–75.
Plateaux, New Zealand. Mar. Micropaleontol. 62, 141–166. Rohling, E.J., Hopmans, E.C., Sinninghe Damsté, J.S., 2006. Water column dynamics
Herguera, J., Berger, W.H., 1991. Paleoproductivity from benthic foraminifera abun- during the last interglacial anoxic event in the Mediterranean (sapropel S5).
dance: glacial to postglacial change in the west-equatorial Pacific. Geology 19, Paleoceanography 21, PA2018. doi:10.1029/2005PA001237.
1173–1176. Rossignol-Strick, M., Paterne, M., 1999. A synthetic pollen record of the eastern
Jorissen, F.J., 1987. The distribution of benthic foraminifera in the Adriatic Sea. Mar. Mediterranean sapropels of the last 1 Ma: implications for the time-scale and
Micropaleontol. 12, 21–48. formation of sapropels. Mar. Geol. 153, 221–237.
Jorissen, F.J., 1999. Benthic foraminiferal successions across Late Quaternary Mediterra- Sabbatini, A., Morigi, C., Ravaioli, M., Negri, A., 2004. Meiofaunal Foraminifera from an
nean sapropels. Mar. Geol. 153, 91–101. abyssal site in the polar front Region (Pacific Sector): size structure, standing stock,
Jorissen, F.J., de Stigter, H.C., Widmark, J.G.V., 1995. A conceptual model explaining taxonomic composition, species diversity and vertical distribution in the sediment.
benthic foraminiferal microhabitats. Mar. Micropaleontol. 26, 3–15. Chem. Ecol. 20, 117–129.
C. Morigi / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 258–271 271

Sangiorgi, F., Dinelli, E., Maffioli, P., Capotondi, L., Giunta, S., Morigi, C., Principato, M.S., Struck, U., Emeis, K.-C., Voß, M., Krom, M.D., Rau, G.H., 2001. Biological productivity
Negri, A., Emeis, K.-C., Corselli, C., 2006. Geochemical and micropaleontological during sapropel S5 formation in the Eastern Mediterranean Sea: evidence from
characterisation of a Mediterranean sapropel S5: a case study from core stable isotopes of nitrogen and carbon. Geoch. Cosmoch. Acta 65 (19), 3249–3266.
BAN89GC09 (south of Crete). Palaeogeogr. Palaeoclimatol. Palaeoecol. 235, 192–207. Sun, X., Corliss, B.H., Brown, C.W., Showers, W.J., 2006. The effect of primary
Schmiedl, G., Mackensen, A., Miller, P.J., 1997. Recent benthic foraminifera from the productivity and seasonality on the distribution of deep-sea benthic foraminifera
eastern South Atlantic Ocean: dependence on food supply and water masses. Mar. in the North Atlantic. Deep-Sea Res. I 53, 28–47.
Micropaleontol. 32, 239–287. Suhr, S.B., Pond, D.W., 2006. Antarctic benthic foraminifera facilitate rapid cycling of
Schmiedl, G., de Bovée, F., Buscail, R., Charrière, B., Hemleben, Ch., Medernach, L., Picon, phytoplankton-derived organic carbon. Deep-Sea Res. II 53, 895–902.
P., 2000. Trophic control of benthic foraminiferal abundance and microhabitat in Suhr, S.B., Pond, D.W., Gooday, A.J., Smith, C.R., 2003. Selective feeding by benthic
the bathyal Gulf of Lions, western Mediterranean Sea. Mar. Micropaleontol. 40, foraminifera on phytodetritus on the western Antarctic Peninsula shelf: evidence
167–188. from fatty acid biomarker analysis. Mar. Ecol., Prog. Ser. 262, 153–162.
Schmiedl, G., Mitschele, A., Beck, S., Emeis, K.-C., Hemleben, C., Schulz, H., Sperling, M., Szarek, R., Nomaki, H., Kitazato, H., 2007. Living deep-sea benthic foraminifera from the
Weldeab, S., 2003. Benthic foraminiferal record of ecosystem variability in the warm and oxygen-depleted environment of the Sulu Sea. Deep-Sea Res. II 54,
Eastern Mediterranean Sea during times of sapropel S5 and S6 deposition. 145–176.
Palaeogeogr. Palaeoclimatol. Palaeoecol. 190, 139–164. Vismara-Schilling, A., Coulbourn, W.T., 1991. Benthic Foraminiferal thanatofacies
Schröder, C.J., Scott, D.B., Medioli, F.S., 1987. Can smaller benthic foraminifera be ignored associated with Late Pleistocene to Holocene anoxic events in the Eastern
in paleoenvironmental analysis? J. Foraminiferal Res. 17 (2), 101–105. Mediterranean Sea. J. Foraminiferal Res. 21 (2), 103–125.
Shackleton, N.J., Sánchez-Goñi, M.F., Pailler, D., Lancelot, Y., 2003. Marine isotope Weldeab, S., Siebel, W., Wehausen, R., Emeis, K.-C., Schmiedl, G., Hemleben, C., 2003.
substage 5e and the Eemian interglacial. Glob. Planet. Change 36, 151–155. Late Pleistocene sedimentation in the Western Mediterranean Sea: implications for
Shepherd, A.S., Rathburn, A.E., Pérez, M.E., 2007. Living foraminiferal assemblages from productivity changes and climatic conditions in the catchment areas. Palaeogeogr.
the Southern California margin: a comparison of the N 150, 63–150, and N63 µm Palaeoclimatol. Palaeoecol. 190, 121–137.
fractions. Mar. Micropaleontol. 65, 54–77. Yanko, V.V., 1990. Stratigraphy and paleogeography of the marine Pleistocene and
Stratford, K., Williams, R.G., Myers, P.G., 2000. Impact of the circulation on sapropel Holocene deposits of the southern seas of the USSR. Mem. Soc. Geol. Ital. 44,
formation in the Eastern Mediterranean. Glob. Biogeochem. Cycles 14, 683–695. 167–187.
Strohle, K., Krom, M.D., 1997. Evidence for the evolution of an oxygen minimum layer at
the beginning of S-1 sapropel deposition in the Eastern Mediterranean. Mar. Geol.
140, 231–236.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Organic geochemistry and paleoenvironment of the Early Eocene “Pesciara di Bolca”


Konservat-Lagerstätte, Italy
Lorenz Schwark a,⁎, Annalisa Ferretti b, Cesare Andrea Papazzoni b, Enrico Trevisani c
a
University of Cologne, Institute for Geology and Mineralogy, Zuelpicher Str. 49a, D-50674 Cologne, Germany
b
Dipartimento di Scienze della Terra, Università di Modena e Reggio Emilia, L.go S. Eufemia 19, I-41100 Modena, Italy
c
Museo di Storia Naturale di Ferrara, Via De Pisis 24, I-44100 Ferrara, Italy

A R T I C L E I N F O A B S T R A C T

Article history: Exceptional preservation of fossils in so-called Konservat-Lagerstätten requires specific depositional regimes
Received 20 December 2007 excluding disturbance of bottom sediments by either wave actions and currents or by benthic fauna. We here
Received in revised form 13 March 2008 describe a depositional model for the Eocene “Pesciara di Bolca” Konservat-Lagerstätte based on sedimentological,
Accepted 14 March 2008
paleoecological, and detailed organic geochemical results. Sediments were deposited in a lagoonal-like basin with
stagnant bottom waters located on an extended carbonate platform that was sheltered from open marine waters
Keywords:
Biomarkers
by a submarine threshold. Run-off from nearby land areas provided nutrients to support an algal community
Highly branched isoprenoid dominated by diatoms. No fossil diatom shells have been identified, but evidence for their presence is given by the
Diatoms high abundance of highly branched isoprenoids in extractable bitumens. Influx of terrigenous organic matter into
Anoxia the lagoon occurred in particular during deposition of the basal fish-bearing level L1. Here not only plant
Fossilization macrofossils, amber, spores and pollen but also the lipid composition indicated notable input of land plants via the
Plattenkalk presence of n-C24 to n-C32 carboxylic acids, long-chain n-alkanes (n-C27, n-C29, n-C31) and angiosperm wax
Lithographic limestone triterpenoids. The redox regime in general was strongly reducing as evidenced by the high concentration of sulfur
Solnhofen
vs. organic carbon, excellent kerogen preservation as shown by high hydrogen indices, and low pristane/phytane
but high phytane/n-C18 ratios. The water column was highly stratified with anoxic saline bottom and fresh surface
waters. Euxinic conditions with free reduced sulfur present in the photic zone could only be detected in sediments
from the L1 horizon via traces of aromatic carotenoids derived from green sulfur bacteria (chlorobiaceae), which
utilize H2S in anoxygenic photosynthesis. The depositional regime is thus comparable to the lithographic
limestones of Solnhofen but based on biomarker evidence lacks the high salinities postulated for the latter.
Biomarker composition indicates that best preservation conditions prevailed in the basal part of the studied
section (0–7 m above datum) but declined upon deposition of the upper part. We interpret the body of the
Pesciara as a parasequence of the 4th order (0.01–0.5 Ma), with the lower part representing a relative sea-level
lowstand and the upper part a relative sea-level highstand.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction While the former leads to an entombment of preferentially benthic


organisms by rapid or catastrophic sedimentation, the latter also
Reconstruction of past environments depends on the interpreta- includes water- or even airborne fauna embedded in sediments
tion of fossil remains, most of which are shelly and thus induce a accumulating in a low hydrodynamic energy regime with reduced
preservation bias in that non-shelly faunas are rarely preserved and water mixing, characteristic of stagnant basins (Seilacher, 1970, 1990).
available for investigation. Under special condition, however, even Konservat-Lagerstätten while depending on exceptional circumstances
soft-bodied organism may be preserved in exceptional detail in so- of accumulation and preservation are not evenly distributed in time
called Fossil-Lagerstätten. These can be divided into concentration and space (Allison and Briggs, 1993). Peaks in outcrop-normalized
deposits and conservation deposits after Seilacher (1970) or Allison occurrence of Konservat-Lagerstätten are coupled to the prevalence of
(1988). The conservation deposits, synonymous to Konservat-Lager- particular environments (Allison, 1988) or evolutionary events (Briggs,
stätten, are related to two major processes: obrution or stagnation. 2003). In the Cambrian, potentially the lack of invertebrate grazers or
deep bioturbators, in the Carboniferous the widespread high pro-
ductivity regimes in coastal delta plains may have favoured preserva-
tion (Briggs, 2003). The Jurassic yields high numbers of Konservat-
⁎ Corresponding author.
E-mail addresses: lorenz.schwark@uni-koeln.de (L. Schwark), ferretti@unimore.it
Lagerstätten due to the extended shelf areas on the NW-margin of the
(A. Ferretti), papazzoni@unimore.it (C.A. Papazzoni), consgeol@comune.fe.it Tethyan Ocean. Here either extensive carbonate platforms with intra-
(E. Trevisani). platform depressions surrounded by microbial or sponge reefs

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.03.009
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 273

restricted circulation and thus favoured deposition of fine-grained sulphurization of biomolecules in an iron-depleted sulphur-enriched
carbonates in Plattenkalk-facies of the Solnhofen type, or transgressive environment.
black shales of the Posidonia Shale type led to exceptional fossil The Pesciara Fossil-Lagerstätte, together with the neighbouring
preservation. The high abundance of small limnic Konservat-Lager- Monte Postale, is the most outstanding fossiliferous outcrop in the
stätten of the Messel type in the Tertiary is dependent on Maar-type surrounding of the Bolca village, about 25 km north-east of Verona
lakes and thus on periods of high volcanic activity. It is thus of major (northern Italy; Fig. 1A). In a few square kilometres several other well-
importance to understand the depositional environment and its known localities are present, such as Monte Purga, Vegroni, and
influence on the modes of fossil preservation when a genetic model of Spilecco, but only for the last one a true exposure is still maintained.
a particular Konservat-Lagerstätte is developed. The Pesciara beds are world-wide known since the 16th century
The Eocene Pesciara di Bolca Konservat-Lagerstätte is characterized (Sorbini, 1972) for their exceptionally preserved fossil fish, but the
by features similar to those found in the Malmian Solnhofen type extraordinary fossil content includes also snakes, birds, insects, land
deposits, mainly the deposition on an extensive carbonate platform plants, amber, algae, rare jellyfish, molluscs, corals, and foraminifera.
with intra-platform depressions or basins, protected from the wash The magnificent quantity and quality of fossil fishes focussed the
and wave action of the open ocean by one or several thresholds attention of the scientists. Amazingly enough, not so much interest
(Barthel et al., 1990; Papazzoni and Trevisani, 2006). The limestones was addressed to fix the geological features (stratigraphy, tectonics,
found in both settings are differentiated into alternation beds of pure sedimentology, geochemistry, etc.) of the area. The age of the Pesciara
micritic limestone (Flinze) and usually much thinner marl beds di Bolca limestone, supposed to be just across the boundary between
(Fäulen). Both settings are assumed to have been shallow water Early and Middle Eocene (Medizza, 1975), has been only very recently
environments, which is characteristic for most Konservat-Lagerstätten, re-assessed as late Early Eocene (Papazzoni and Trevisani, 2006).
though deep-water settings (N200 m depth) are reported for the La It is the aim of the current study to investigate the organic matter
Voulte Fossil-Lagerstätte, an environment characterised by sponge occurring in samples from the Pesciara Konservat-Lagerstätte in order
reefs equivalent to the Solnhofen type deposits (Charbonnier et al., to improve the understanding of its depositional environment and to
2007). The environment of deposition must have been anoxic and compare this site with other Konservat-Lagerstätten of similar origin.
eventually euxinic due to the lack of bottom dwellers, except for In this respect the bulk kerogen and the solvent-extractable bitumen
washed in organisms or rare currents marks. Anoxygenic conditions fraction will be analyzed and results compared with sedimentological
enhance the preservation of organic matter primarily produced within and paleocological features available for the same sample set (Papaz-
the water column by autotrophic organisms and also that of floating or zoni and Trevisani, 2006).
benthic microbial mats, these being formed by autotrophic or hetero-
trophic bacteria. 2. Geological setting of the Pesciara Konservat-Lagerstätte
The organic matter content of the Solnhofen and the Pesciara di
Bolca limestones, however, is rather low with TOC b0.5% (Barthel et al., The Pesciara beds appear in the field as a white “spot” (a few
1990; Schwark et al., 1998; Papazzoni et al., 2007), whereas TOC values hundred square metres) of limestones surrounded by volcanic rocks
of up to 15% were described for the marl intervals (Schwark et al.,1998). on the right side of Val del Fiume, about 2 km north-east of Bolca
Similarly low TOC values were reported for the Plattenkalke from (Fig. 1A). The strata, plunging south-easterly, are less than 20 m thick.
southern Germany (0.4–0.9%; Hückel, 1974), the Lower Cretaceous La All the calcareous rocks of Early–Middle Eocene age in the western
Huérguina Limestone Formation of Las Hoyas, Spain (0.37–1.34%; Veneto area are collectively known as “Calcari nummulitici” (nummu-
Buatois et al., 2000) or the Late Cretaceous Hvar Facies of Hvar, Croatia litic limestones). Indeed, this name refers to different local lithos-
(0.73–1.23%; Swinburne and Hemleben, 1994). In accordance with tratigraphical units sharing a roughly similar fossil content. Even the
these data, exceptional fossilization in anoxic conditions does not Pesciara limestones do not belong to any formalized unit. The lack of
appear to guarantee high organic matter contents. formalized units hinders so far any possible reconstruction of the
The low concentration of organic matter until now has restricted development of the carbonate platform (“Lessini Shelf” sensu Bosellini,
the application of organic geochemical techniques to the study of the 1989) during this time span.
Bolca and Solnhofen deposits. Solnhofen type deposits have been A possible correlation exists between the Pesciara laminated
preliminary investigated by Keupp et al. (1993) and Hefter et al. (1993) limestones and similar rocks cropping out in the nearby Monte Postale
using outcrop samples of mainly spongiolitic microbialites and in more (about 300 m away). Unluckily, due to the widespread occurrence of
detail by Schwark et al. (1998) using drill core samples from well volcanic and volcanoclastic rocks and to the tectonic displacement of
Stoerzelmuehle. Several Upper Kimmerdigian and Lower Tithonian the sedimentary beds, the relationships between these localities are
strata in Franconia yield lithographic limestones and stratigraphic still unclear. This will be matter for a forthcoming paper.
misplacement of fossils has occurred frequently in the past (Fürsich Bolca lays East of the Castelvero Line, a tectonic element running
et al., 2007). It may thus be preferable to refer to the “Solnhofen type” NNW–SSE on the right side of the Val d'Alpone. This fault complex
rather than to the Solnhofen location. The core investigated by determined and controlled the volcano-tectonic and paleoenviron-
Schwark et al. (1998) derived from the Rennertshofen Trough, which mental evolution of the Lessini area during the Late Paleocene–Middle
is separated from the Solnhofen Trough by a sponge–microbial reef Eocene. Then, the Castelvero Line acted as a tectonic threshold sepa-
complex and thus represents a slightly different facies. The molecular rating a western sector with thin and discontinuous volcanic products
geochemical data for the Solnhofen Trough indicated a high variability in from an eastern sector where the major part of volcanic (basaltic) rocks
organic matter sources (Hefter et al., 1993; Keupp et al., 1993) including accumulated during this time span.
molecular fossils derived from eubacteria, cyanobacteria, archaea and During the same interval, there were several pulses of activity
land plants. Molecular indicators of enhanced salinity as postulated for involving the Castelvero Line, generating a stepwise subsidence of the
the Solnhofen depositional environment (Barthel et al., 1990) could not eastern sector, which became a tectonic low known as Alpone-Agno
be found by Hefter et al. (1993) or Keupp et al. (1993). For the semigraben (Barbieri et al., 1982, 1991). Here volcanic rocks accumu-
Rennertshofen Trough adjacent to the classical Solnhofen location lated, intercalated with marine carbonates during the volcano-
(distance about 40 km) biomarker results indicated enhanced salinity tectonic stasis (Barbieri, 1972; De Zanche and Conterno, 1972). The
and most likely a salinity/density-stratified water column in the extensive tectonic movements, mainly documented near the Castel-
platform depression surrounded by sponge/microbial reefs. Biomarker vero Line, were discontinuous and coincident with the main volcanic
data for samples from well Stoerzelmuehle revealed a high variability in activity. Six phases were recognized by Barbieri et al. (1991) between
organic matter input from land and marine sources and intensive the Late Paleocene and the Middle Eocene.
274 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

Fig. 1. Location map (A) and stratigraphic column (B) of the Pesciara Section (simplified after Papazzoni and Trevisani, 2006).

The Monte Postale and Pesciara carbonate deposition probably Four main microfacies types were recognized by Papazzoni and
occurred during the interval between the third and the fourth volcanic Trevisani (2006). The M microfacies characterizes autochthonous
phase of Barbieri et al. (1991). micritic limestones (levels L1 to L5) rich in fish and plants but with few
The fair continuity of the carbonate deposits East and West of the recognizable bioclasts. A further subdivision in three subunits (M1, M2
Castelvero Line suggests that the tectonic accommodation repeatedly and M3 microfacies) was attempted on the basis of the presence of
produced East of the fault was quickly filled in by the volcanic products. black laminae (M1), white sparitic laminae (M2) or of the absence of
Within the Alpone-Agno semigraben, the sediments of Paleocene– any lamination (M3). The F microfacies typically consists of biocalcar-
Eocene age were subjected to brittle deformation, subdivided into enite–biocalcirudites containing abundant resedimented benthic
separate blocks by plane and subparallel faults in a generally extensive organisms, mainly foraminifera and molluscs, indicative of a very
setting. The blocks were tilted and moved with a domino-style shallow environment. This facies often occurs in alternation with the
arrangement (Barbieri and Zampieri, 1992; Zampieri, 1992, 1995). M microfacies. The B microfacies represents a calcareous-siliceous
The Bolca area belongs to a tectonic low bordered on the West by breccia with extraclasts present in the lower part of the section.
the Castelvero Line faults, running NNW–SSE, and on the East by the Finally, the N microfacies, limited to a single submittal sample, typifies
Monte Postale fault (Barbieri and Medizza, 1969; Dal Degan and biocalcarenites–biocalcirudites with an autochthonous association of
Barbieri, 2003), a subvertical line nearly parallel to the Castelvero Line. nummulites and assilinas.
This structure probably created the space for the accommodation of Papazzoni and Trevisani (2006) recently proposed a depositional
the carbonate sediments during the Eocene. model for the Pesciara Section, referring to a shallow basin with
restricted circulation (“lagoon”) and anoxic conditions prevailing in
3. Investigated material stagnant bottom waters.
The material used in this work comes from the recent sampling of
Most of the spectacular biota recovered in the classic Pesciara Papazzoni and Trevisani (2006) both on the surface and into tunnels
Section of Bolca derives from fine-grained and evenly laminated opened to extract the fossil fishes. Either laminated or the coarse-
limestones, up to 1 m thick, organized in five distinct major levels (L1 grained calcareous levels were selected (Table 1). As regards the
to L5; Papazzoni and Trevisani, 2006). These productive layers are former, levels L1 (samples 0201 and 0202), L2 (sample 0207) and L3
intercalated with storm-induced coarse-grained limestones rich in (sample 0214) have been analysed. They characterize the M micro-
molluscs and foraminifera, but devoid of any fish remnants (Fig. 1B). facies. Within the coarse-grained beds (F microfacies), we used
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 275

Table 1
Stratigraphic location and main features of levels investigated for geochemistry

Sample no. Short description Position within or m above datum Microfacies (⁎) Notes
relative to fish levels
0223 Wackestone with small foraminifera. Above L4 17.00 M3 –
0220 Packstone–rudstone with bivalves and alveolinas. Above L4 15.10 F –
0218 Faintly-laminated, marly wackestone. Under L4 (see note) 13.75 M3 Nucleus of the slump involving L4.
0216 Non-laminated, marly wackestone. Some coarser laminae Between L3 and L4 10.15 M3-F –
with alveolinas intercalated.
0215 Non-laminated to faintly-laminated, marly wackestone. Between L3 and L4 9.25 M3 In vicinity to a basaltic dyke.
0214 Laminated limestone. Some coarser laminae with miliolids L3 7.85 M2 Fish are often deformed in this level.
intercalated.
0212 Wackestone with miliolids and alveolinas, in places faintly Between L2 and L3 7.00 F –
laminated.
0207 Laminated limestone. L2 6.35 M1 –
0206 Grainstone with abundant miliolids. Intercalated within L2 5.90 F –
0204 Packstone with alveolinas. Between L1 and L2 3.65 F –
0203 Bioclastic, brecciated packstone with alveolinas. Many clasts Just over the top of L1 2.70 B –
are silicified.
0202 Clayey-silty marl, evenly laminated. L1 1.00 M1 –
0201 Evenly-laminated limestone. L1 0.50 M1 –

⁎ refers to microfacies types described by Papazzoni and Trevisani (2006) and briefly summarized in the text.

material from the massive biocalcarenite–biocalcirudites (samples and extract yields determined gravimetrically. The low extract yields
0204, 0212 and 0220), from the centimetre-thick bioclastic levels of commonly 0.5 to 2 mg prevented sophisticated compound class
(samples 0206 and 0216) and from those levels revealing either no separation by MPLC as commonly applied (Schwark et al., 1998). Total
lamination or a faint lamination (samples 0215, 0218 and 0223). extracts were thus derivatized with bis-silyl-trifluoracetamid (BSTFA)
Finally, even the basal calcareous-siliceous breccia rich in benthic to make polar functionalized lipids amenable to GC/MS analyses.
fossils (sample 0203; B microfacies) has been investigated. GC/MS analysis of derivatized extracts was carried out using a HP
5890 Series II gas chromatograph coupled to a HP 5989 single
3.1. Analytical methods quadrupole mass spectrometer. A 50 m long fused silica column (HP 5)
with an internal diameter of 0.25 mm, coated with 5% chemically
For all sampled levels, total carbon, nitrogen content, organic carbon bonded phenyl-polysiloxane (0.25 µm) was used with helium as
content and stable isotopic analyses of organic C were determined in carrier gas at a constant flow rate of 1.0 ml/min. Injection was done in
Bologna, Italy. After this first set of analyses, material from the same the splitless mode at a temperature of 300 °C. Following an isothermal
levels was prepared for kerogen analyses and specific biomarker iden- period of 2 min, the GC oven temperature was programmed from 70 to
tification in Cologne, Germany. 140 °C at a rate of 10 °C/min followed by a second gradient from 140 to
Total carbon and nitrogen were determined using a FISONS NA2000 320 °C at 3 °C/min. The mass spectrometer operated in EI mode at
Elemental Analyzer (EA). The organic carbon contents were determined 70 eV with a scan range from m/z = 50 to 600 for identification of
using the same instrument after removal of the carbonate fraction by individual compounds. Data acquisition and processing was per-
dissolution in 1.5 normal HCl. Stable isotopic analyses of organic C were formed using a HP MS-ChemStation data system. Peak identification
carried out by using a FINNIGAN Delta Plus mass spectrometer, which was based on retention times and on comparison of mass spectra with
was directly coupled to the FISONS NA2000 EA by means of a CONFLO those of the system library and published spectra.
interface. The IAEA standard NBS19 was used as calibration material for
carbon. Uncertainties are usually lower than ±0.2‰, as determined from 4. Results
routine replicate measurements of a reference sample. Stable isotopic
data are expressed in the conventional delta (δ) notation in which the 4.1. Carbonate content
13 12
C/ C isotopic ratios are reported relative to the international VPDB
standard. The amount of carbonate was independently assessed by the loss
Homogenized, finely crushed and ground samples were decalcified upon HCl treatment and by the difference between total carbon (TC) and
with 10% HCl and washed till neutral. The carbonate content was organic carbon (TOC) yielding the total inorganic carbon (TIC). Under the
calculated by difference between sample material subjected to HCl assumption that the carbonate phase is calcite, TIC ⁎ 8.33 yields the
treatment and residual fraction remaining thereafter. The decalcified carbonate content. Elemental data generated in Italy (CNR-ISMAR in
sample powder was then subjected to elemental analysis and Rock-Eval Bologna) are listed in Table 2 and agree well (R2 = 0.9877) with the
analysis. Elemental analysis was conducted for total organic carbon of carbonate content determined by HCl dissolution in Cologne (Table 3)
the decalcified sample TOCcf and total sulfur of the decalcified samples though the carbonate determination using the TIC systematically
TScf using a LECO CS225 Elemental analyzer. overestimates the actual concentration. This must be due to calibration
Rock-Eval analysis was performed according to the method problems with carbonate rich samples as no indication of a different
described by Espitalié et al. (1977) and Bordenave et al. (1993) with carbonate phase and thus lower stoichometry factors is evident.
a VINCI Rock-Eval-II Analyser. S1, S2 and Tmax values were measured Except for sample 0203, which is a silicified breccia with a
and corresponding HI, OI and PI values calculated. For details of carbonate content of less than 15%, most other samples contain N90%
procedure and interpretation see Lüniger and Schwark (2002). carbonate, as also confirmed by X-ray analyses. Lower carbonate
Bitumen was extracted from bulk rock by accelerated solvent values of 82 and 84.5% were encountered in samples 0202 and 0206,
extraction (DIONEX ASE 200) with dichloromethane (DCM) at 75 and respectively. Sample 0202 is a dark-coloured marl bed from the main
120 °C and 50 bar pressure, whereafter the extracts were combined. fish-bearing level L1 in the Pesciara Section (Papazzoni and Trevisani,
Elemental sulfur was precipitated from the raw extract by addition of 2006), where the non-carbonate fraction is assumed to be clay, pyrite
activated copper foil. After filtration samples were taken to dryness and abundant organic matter (Table 2). Sample 0206 does not contain
276 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

Table 2
Organic geochemical data for raw samples from the Eocene Pesciara Section

Sample Lithology TN (%) TC TOC (%) TIC (%) δ13C (‰ VPDB) TOC/TN Extract (ppm) Extract (mg/gTOC) Carbonate (%)
0223 Calcarenite 0.010 12.98 0.46 12.52 −25.6 45.3 64 14.0 104.3
0220 Calcarenite 0.005 13.02 0.19 12.83 −26.0 34.2 58 31.0 106.9
0218 Calcarenite 0.009 13.13 0.36 12.77 −26.0 41.2 50 13.8 106.3
0216 Bioclasts 0.009 13.05 0.33 12.72 −26.3 36.9 81 24.3 105.9
0215 Calcarenite 0.014 12.27 0.55 11.72 −24.0 39.4 72 13.1 97.7
0214 Laminated micrite 0.007 12.98 0.35 12.63 −25.5 47.5 68 19.4 105.2
0212 Calcarenite 0.006 12.84 0.24 12.60 −26.3 37.4 24 10.1 105.0
0207 Laminated micrite 0.011 12.78 0.49 12.29 −24.7 45.4 85 17.4 102.4
0206 Calcarenite layer 0.005 11.04 0.16 10.88 −25.0 32.5 14 8.8 90.7
0204 Calcarenite 0.004 12.78 0.16 12.63 −26.7 37.5 3 1.8 105.2
0203 Siliceous breccia 0.005 1.30 0.21 1.09 −26.3 44.3 32 15.5 9.1
0202 Laminated micrite 0.195 18.04 8.60 9.44 −24.2 44.0 1486 17.3 78.6
0201 Laminated micrite 0.012 12.43 0.48 11.95 −25.6 38.6 85 17.6 99.5
Average 0.023 12.20 0.97 11.24 −25.6 40.3 163 15.7 93.6

abundant clay and organic matter indicating that most of the non- 4.3. Organic carbon isotopic composition
carbonate fraction is due to silicification.
The δ13C isotopic composition of the Pesciara samples ranges
4.2. Organic carbon, total nitrogen, and total sulfur content between −24.0‰ and −26.7‰, thus indicating an origin from C3-
photosynthesis by aquatic or terrestrial organisms. Values of bioclastic
The organic carbon contents of samples from the Pesciara Section limestones (mean −26.1‰) are notably lighter than those of the
are low, with values between 0.16 and 0.55% TOC, except for sample productive layers (mean −25.2‰).
0202 showing an exceptional value of 8.60% TOC. The total nitrogen
concentrations closely follow the TOC concentrations (Table 2). The C/ 4.4. Rock-Eval analyses
N ratios of the organic matter varied between 33 and 48, with a mean
42.8 value for productive layers and a mean 37.2 value for bioclastic Investigation of sedimentary organic matter by Rock-Eval provides
layers (Table 2). For samples of this age and diagenetic stage this does information on the type of organic matter accumulated, the thermal
not allow for a differentiation of marine versus terrigenous matter maturity stage of the kerogen, and indicates the presence of migrated
prevalence (Tyson, 1995). The low TOC values preclude a detailed bitumen or impregnation (Espitalié et al., 1977; Peters, 1986;
interpretation of environmental change across the profile investi- Bordenave et al., 1993; Lüniger and Schwark, 2002). Interpretation
gated. This compares well to the Malmian carbonates of the of Rock-Eval data is only recommended, if the TOC value of the
Franconian Alb in Germany, where very low TOC concentrations sediments to be investigated is N0.5% TOCcf (Espitalié et al., 1977;
prevailed in the Solnhofen location but enhanced TOC contents were Peters, 1986; Bordenave et al., 1993; Langford and Blanc-Valleron,
noted in samples from well Stoerzelmuehle. The low concentrations of 1990). Insufficient amount of organic matter leads to erroneous results
TOC in the finely laminated carbonates from Solnhofen are assumed to due to unpredictable mineral matrix effects, low signal intensities and
be at least partially due to post-sedimentary oxidation and loss of associated unacceptably low signal to noise ratios. Given these
organic matter (Barthel et al., 1990; Schwark et al., 1998). In analogy it limitations, only but one of the samples from the Bolca location
may be assumed that the fish-bearing levels L2 to L4 originally would have been suitable for analyses. The decalcified samples
contained higher concentrations of organic matter but lost a notable subjected to the Rock-Eval analyses with the exception of the breccia
proportion upon later weathering. This is supported by the low degree (Table 3) all fulfill the requirements for Rock-Eval analyses, given that
of correlation between TOCcf and TScf values of decalcified samples TOC concentrations of the residual fractions average 12.2%. The
(Table 3). The sulfur content of decalcified samples is very high with kerogen concentration in the residual fraction is high, with an average
values up to 56% in sample 0216, due to intensive pyritisation. The fact of 55 mg hydrocarbons per gram of rock (Table 3). Hydrogen index
that at least some sedimentary intervals of both Konservat-Lager- (HI) values for the Pesciara Section sediments vary between 237 and
stätten still bear relicts of concentrations of primary organic matter 626 (mg HC/g TOC) as shown in Table 3. The low HI values of the
content indicates deposition under anoxic stagnant water conditions. breccia sample are due to the lack of indigenous kerogen, most of the

Table 3
Organic geochemical data for decalcified samples from the Eocene Pesciara Section

Sample Lithology HCl soluble (%) TOC (%) S (%) Tmax (°C) S1 S2 S3 PI HI OI
residual residual (mg HC/g rock) (mg HC/g rock) (mg CO2/g rock) (mg HC/g TOC) (mg CO2/g TOC)
0223 Calcarenite 98.4 24.80 12.45 407 8.1 108.7 10.0 0.07 438 40
0220 Calcarenite 99.4 20.62 14.83 405 3.8 72.36 8.5 0.05 351 41
0218 Calcarenite 98.8 18.51 19.13 406 7.4 93.42 6.4 0.07 505 34
0216 Bioclasts 97.8 6.67 56.76 409 1.8 23.60 3.4 0.07 354 51
0215 Calcarenite 97.9 12.34 19.36 406 6.8 39.33 2.5 0.15 319 20
0214 Laminated micrite 98.9 17.35 10.29 407 8.6 108.6 3.1 0.07 626 18
0212 Calcarenite 99.5 11.32 1.44 418 2.0 41.36 3.6 0.05 365 32
0207 Laminated micrite 98.6 21.30 3.33 412 10.1 120.1 3.7 0.08 564 17
0206 Calcarenite layer 84.5 0.91 0.33 415 0.2 2.16 0.5 0.08 237 55
0204 Calcarenite 98.7 1.60 0.10 424 0.6 5.55 1.3 0.09 346 82
0203 Siliceous breccia 14.7 0.16 0.00 384 0.1 0.03 0.1 0.79 – 32
0202 Laminated micrite 82.0 16.35 8.60 408 9.5 81.00 5.3 0.11 495 32
0201 Laminated micrite 93.3 6.99 12.46 400 4.8 25.60 1.2 0.16 366 17
Average 89.4 12.22 12.24 408 4.9 55.53 3.8 0.14 414 36
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 277

organic matter present being a product of impregnation from adjacent


strata. This is to be recognized by the exceptionally high production
index (PI) reaching a value of 0.79 (Table 3). In general the PI values are
very low (b0.1) or low in samples 0201, 0202, and 0215 (b0.15), which
is indicative of the low thermal maturity of the organic matter. This is
supported by the low Tmax values of an average 408 °C, except for the
breccia sample where the low TOC content leads to erroneous Tmax
determination. Higher Tmax values are encountered in samples 0204
to 0212, which might indicate the presence of more refractory organic
matter. The Rock-Eval data show no evidence for enhanced thermal
maturation of the organic matter due to the volcanics emplaced
within the Pesciara Section.
The oxygen content of the organic matter (OI) serves as a proxy for
the amount of terrigenous matter being present. The OI values of
samples from the Pesciara site average 36 (mg CO2/gTOC) with
elevated values noted for samples 0204 and 0206 that already showed
higher proportions of non-marine refractory material by their
corresponding Tmax values. A higher OI value also occurs with the
bioclastic calcarenite of sample 0216, remarkable by its exceptional
degree of pyritization.
HI versus OI values or HI versus Tmax values allow to classify the
kerogen in the Pesciara sediments as marine type II of low thermal
maturity (Fig. 2). The mineral matrix effects (Langford and Blanc-
Valleron, 1990; Lüniger and Schwark, 2002) affecting raw samples to a
high degree are almost absent in the decalcified Pesciara samples. This
leads to a regression line with an intercept close to the origin (Fig. 2)
indicative of no retardation of the pyrolizate. Rock-Eval data on
decalcified limestones clearly indicate that despite of the low
concentrations of total organic matter, the quality of preservation
was high in the Pesciara basin and favored the development of a true
Konservat-Lagerstätte.

4.5. Bitumen content

The amount of extractable bitumen is closely correlated with the


TOC values (R2 = 0.988) and varies between 3 and 1486 ppm (Table 2). If
normalized to TOC content the low values of on average 15.6 (mg Ext/
gTOC) indicate an immature stage of kerogen evolution. This is in
agreement with PI and Tmax values determined in Rock-Eval analyses.
The low absolute amounts of bitumen prevented a compound class
separation by liquid chromatography and thus total extracts were
analyzed by GC/MS for the presence of source and environment
characteristic biomarker molecules. In recognition of the permeability
of the carbonate rocks and the mobility of bituminous fluids in the pore
space of the subsurface care has to be taken to make sure that the
bitumen is of autochthonous nature when organic facies interpretation
is envisaged (Peters et al., 2005). The low PI values of most samples and
the high variability in extract yields and composition argue against an
impregnation of the pore space by a homogeneous allochthonous fluid.
Except for few samples (see below) the bitumen composition will thus
be indicative of environmental change across the Pesciara profile.

4.6. Molecular composition

Organisms biosynthesize a large variety of different molecules in


response to the environmental demands prevailing. Many of the
characteristic biomolecules can be preserved with minor or no alteration

Fig. 2. Rock-Eval data for source and maturity identification. A) A pseudo van Krevelen
diagram depicts samples as type II kerogen of low maturity, with some degradative
alteration and admixture of terrigenous material. B) The HI versus Tmax diagram
confirms a type II kerogen of low maturity that has been affected by secondary
alteration. C) An improved kerogen classification diagram minimized the mineral
matrix effects yielding an averaged HI value of 495 (mg HC/g TOC) calculated from the
slope of the regression line.
278 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

Table 4 ubiquitous isoprenoidal alkanes phytane, pristane, norpristane and


Selected biomarker data for samples from the Eocene Pesciara Section smaller analogues, a suite of saturated steranes with different degree of
Sample Lithology Pri/Phy Phy/ Short/long n-C16/ HBI20/ HBI25/29 alkylation at the C24-position and a series of hopanes and hopenes.
n-C18 n-alkanes n-C17 n-C17 Functionalized lipids are mainly represented by carboxylic acids,
0223 Calcarenite 0.87 0.55 4.14 1.68 0.39 2.80 whereas alcohols were not determined in notable concentrations. The
0220 Calcarenite 0.95 0.39 5.39 2.38 0.26 4.05 fatty acids in all but one sample are exclusively medium-chain
0218 Calcarenite 0.68 0.82 2.78 1.61 0.48 1.98
homologues in the C14 to C18 range, predominantly occurring as n-
0216 Bioclasts 0.91 0.52 7.89 2.29 0.48 2.33
0215 Calcarenite 0.96 0.75 2.01 1.94 0.77 1.97 fatty acids with subordinate amounts of branched and unsaturated
0214 Laminated micrite 0.87 0.58 6.14 1.60 0.51 2.35 fatty acids (Fig. 4).
0212 Calcarenite 0.86 0.51 3.77 1.30 0.38 2.13 The only exception from this pattern is given by sample 0202,
0207 Laminated micrite 0.65 1.73 1.15 0.95 2.86 4.44 which is the most TOC-rich and unaltered. Here a suite of long-chain
0206 Calcarenite layer 0.64 0.82 1.54 0.82 2.73 8.71
0204 Calcarenite 0.90 0.58 3.44 0.88 0.31 0.79
fatty acids extending up to the n-C32 homologue is present (Fig. 4). The
0203 Siliceous breccia 0.77 0.59 2.10 0.82 0.03 2.84 carbon preference index of these n-carboxylic acids attests to the low
0202 Laminated micrite 0.55 3.25 0.73 0.81 2.04 5.86 thermal maturity of the extractable bitumen and clearly indicates an
0201 Laminated micrite 0.65 1.57 1.99 1.33 1.10 3.44 origin from cuticular waxes of land plants.
Average 0.87 0.59 4.59 1.42 0.95 2.52
Cyclic biomarkers occurred in low concentrations close to the
detection/identification limit in most samples. The cyclic biomarker
distribution is dominated by triterpenoidal steranes, hopanes and
hopenes. Functionalized triterpenoids are restricted to hopanoic acids.
in sediments and these geomolecules provide evidence on the Despite of the low maturity of the samples no sterenes, steradienes or
organisms that once thrived in the environment of deposition. their respective rearranged counterparts were detected. The hopanoid
Molecules that are indicative of a specific biological precursor are called distribution was dominated by ββ-hopanes and contained diploptene
biomarkers or chemofossils and serve in the reconstruction of organic but no hop(17,21)enes usually abundant in low maturity samples of
facies (Table 4). For a more detailed introduction into the principles of comparable organic facies, e.g. the Solnhofen type limestone (Schwark
organic geochemistry the unfamiliar reader is referred to Peters et al. et al., 1998). Though not quantified in detail due to the low extract
(2005), Killops and Killops (2005), and Brocks and Summons (2007). yields, the proportion of steroids outcompetes that of hopanoids
The distribution of biological markers found in the Pesciara samples revealing a dominance of algal over eubacterial/cyanobacterial lipid
is illustrated by four typical and representative gas chromatograms sources. The sterane distribution shows higher abundances of the C27
shown in Fig. 3. The hydrocarbon components identified comprise a and C29 analogues. The former represent an algal input, in particular
series of n-alkanes ranging from n-C12 to n-C34, a series of highly from rhodophytes, the latter may derive from green or brown algae but
branched isoprenoids (HBI) with 20, 25, 29 and 30 carbon atoms, the also from land plants (Killops and Killops, 2005; Peters et al., 2005). The

Fig. 3. GC/MS mass fragmentograms (m/z 85) of four characteristic samples from the Pesciara Section, revealing the distribution of aliphatic hydrocarbons.
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 279

low concentrations of C28 steroids stand in contrast to their prefe- tion of elution time and mass spectral fragmentation pattern to occur
rential biosynthesis by diatoms and prymnesiophytes (Peters et al., in the Pesciara extracts in notable concentrations (Fig. 3), so far has not
2005), both algal classes assumed to be abundant in the Pesciara basin. been reported.
Sample 0202 from the base of the section, giving the highest Short-chain n-alkanes preferentially derive from aquatic primary
bitumen yield was characterized as having received a notable input of producers, in particular algae and cyanophytes but may also be
land plant derived alkyl lipids and in addition also showed triterpe- attributed to heterotrophic bacteria (Peters et al., 2005). Their do-
noids of the oleanene type, characteristic of angiosperm plants. No minance in all samples documents the predominantly marine organic
aliphatic or aromatic diterpenoids, indicative of gymnosperm input matter sources in the Pesciara Section. Except for sample 0202 all
into the Pesciara basin were detected in the total extract fractions. carboxylic acid fractions are dominated by saturated short-chain fatty
acids (Fig. 4) again indicative of marine autotrophs but also hetero-
5. Discussion trophic bacteria. Iso- and antiso-C15 and C17 fatty acids, which are
indicative of bacterial sources, e.g. sulfate reducers, occur in minute
The organic geochemical data determined for the bulk kerogen and amounts only in sample 0202 and generally lack in the remaining
for the extractable biomarkers allows addressing the following issues: samples thus emphasizing a surprisingly low abundance of bacterial
i) identification of marine algal and microbial groups contributing to biomass, as already indicated by the low hopanoid versus steroid
the organic matter, ii) identification of terrigenous input into the concentrations.
Pesciara basin, iii) the redox potential prevailing in the sediments, iv)
assessment of the salinity in the Pesciara basin and v) recognition of 5.2. Terrigenous organic matter
stratigraphic variability in environmental conditions.
The bulk geochemical data revealed a low input of terrigenous
5.1. Marine organic matter organic matter, which is in agreement with the overwhelmingly pure
carbonate lithology that does not allow for large quantities of
The biomarker distribution of most samples is dominated by HBI terrigenous influx. The HI and OI values from Rock-Eval analyses
compounds that are biosynthesized by a variety of diatoms (Yon et al., indicate a marine organic matter with variable degree of oxidative
1982; Rowland and Robson, 1990; Sinninghe Damsté et al., 2004). The degradation. The lipid composition, however, indicates terrigenous
relative distribution of HBI in diatoms is dependent on species as well influx, especially in the lower section and here, in particular in sample
as growing conditions but a differentiation of diatom type or 0202 (Figs. 3 and 4). Long chain n-alkanes and carboxylic acids are well-
environment based on relative abundance of individual HBI is not known and reliable indicators of land plant waxes (Killops and Killops,
yet possible (Peters et al., 2005). Whereas C20, C25 and C30 HBI have 2005) and are clearly recognizable in sample 0202. Plant waxes in
been described frequently and in rare occasions C35 HBI have been contrast to woody plant material can reach the marine or limnic
found (Peters et al., 2005), a C29 analogue as inferred from interpreta- environment by eolian transport (Gagosian and Peltzer, 1986; Gagosian

Fig. 4. GC/MS mass fragmentograms (m/z 117) of four characteristic samples from the Pesciara Section, revealing the distribution of carboxylic acids.
280 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

et al., 1987; Lüniger and Schwark, 2002). Therefore, they are common in indicate eolian influx, whereas the macrofossil evidence and the high
marine settings receiving little or no fluviatile input (Rontani and abundance of terrigenous lipids in sample 0202 may originate from
Volkman, 2005). The wax lipids then contribute only a minor proportion fluviatile transportation.
to the bulk kerogen but occur exclusively in solvent-extractable form.
Because of this and because of their recalcitrance (Peters et al., 2005; 5.3. Redox regime
Marschner et al., 2008) they are overrepresented in the solvent extracts
of the kerogen lean-samples. The cyclic biomarkers indicate input of Preservation of biomarker lipids and primary composition of
angiosperm lipids whereas gymnosperm lipids are not detectable, but biomolecules synthesized by organisms occupying a given environment
Trevisani et al. (2005, Table 2) reported pollens of Gymmosperms, highly depend on the oxygen availability or redox regime. A classical
Auracariaceae (genus Araucariacites), plus some spores of Bryophyta indicator of redox conditions is given by the breakdown of chlorophyll,
(genus Stereisporites) and Pteridophyta (genera Gleichenidiites, Leiotri- ubiquitous in phototrophs, leading to either phytane under reducing or
letes, Cicatricosisporites, Echinatisporis), all from L1. This discrepancy pristane under oxidizing conditions (Powell and McKirdy, 1973; Didyk
may be explained by the low amount of free lipids in spores and pollen, et al., 1978). The pristane/phytane ratio though it may be affected by
which thus do not contribute to the biomarker yield of the sediments. various factors (ten Haven et al., 1985; Frimmel et al., 2004; Peters et al.,
Wind transported plant wax rich in angiosperm triterpenoids, however, 2005) indicates reducing conditions by values b1.0. According to this
contribute significantly to the free lipids but do not leave microscopically threshold value all samples from the Pesciara site were deposited under
identifiable objects. A terrigenous influx agrees well with plant reducing conditions, with lower Eh values in the lower and higher Eh
macrofossils (Massalongo, 1859a,b; Papazzoni and Trevisani, 2006) values in the upper part of the section (Fig. 5). The molecular redox
though the biomarkers in the pure carbonates may preferentially indication corresponds to the high preservation potential documented

Fig. 5. Trends in biomarker distribution along the Pesciara profile: A) pristane/phytane ratio, B) phytane/n-C18 ratio, C) hydrogen index HI (mg HC/g TOC), D) δ13Corg (‰), E) short/
long-chain n-alkanes, F) n-C16/n-C17 ratio, G) HBI20/n-C17 ratio, H) HBI25/HBI29 ratio. Open symbols represent samples 0203 and 0204, both affected by impregnation.
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 281

by the kerogen Rock-Eval data though the latter are not suitable for hypersalinity empirically a value of 0.3 has been established (ten Haven
differentiation of redox gradients across the profile (Fig. 5). In a similar et al., 1985), which is not reached in the Pesciara Section. In hypersaline
fashion, the relative proportions of phytane and octadecane find environments the diagenesis of steroids is supposed to follow a different
application as redox proxy. The higher phytane/n-C18 ratio in the route, leading to high concentrations of ββ-steranes at low maturity
lower part of the Pesciara Section and the trend to lower values in the levels (de Leeuw and Sinninghe Damsté, 1990). The sterane distribution
upper part confirm the Eh assessment based on the pristane/phytane of the Pesciara samples is dominated by C27 to C29ααα steranes with the
ratio. The environment of deposition is thus interpreted as having been biological 24R configuration, thus showing no indication of enhanced
severely oxygen depleted (anoxic) though most likely not enriched with salinity. The distribution of alkylated chromans has been applied for
reduced sulphur species within the water column (euxinic). This is identification of paleosalinity levels (Sinninghe Damsté et al., 1987, 1993;
inferred from the lack of sulphurized biomarkers, commonly formed Schwark and Püttmann, 1990; Schwark et al., 1998) but has been
during early stage of diagenesis in comparable settings, e.g. the questioned by Li et al. (1995). No alkylated chromans were detected in the
Solnhofen type deposits (Schwark et al., 1998) and from the absence of Pesciara sediments and thus no salinity assessment is possible based on
aromatic carotenoids and their derivatives. In shallow stratified water their relative distribution. In hypersaline sulfur-enriched environments
columns with euxinic bottom water microbial communities bearing early incorporation of reduced sulfur species is common and allows for
green sulphur bacteria (chloriobiaceae) thrive within the photic zone salinity discrimination based on a variety of thiophene and thiolane
(Summons and Powell, 1986; Schwark and Püttmann, 1990; Grice et al., compounds (Sinninghe Damsté et al., 1989, 1990; Schwark et al., 1998). In
1996a). Whereas aryl isoprenoids are common and diagnostic in the the Pesciara samples, no organic sulfur compounds were detected and
Fossil-Lagerstätte of the Toarcian Posidonia Shale (Schwark and Frimmel, thus no paleosalinity information could be derived. In summary all
2004), they were not detected in the Jurassic limestones of the common molecular paleosalinity indicators either showed no values
Solnhofen type (Schwark et al., 1998) and in the Pesciara Section indicative of enhanced salinity or the indicator compounds were not
occurred in trace amounts exclusively in sample 0202. Chlorobiaceae present. This multiple evidence strongly implies that the salinity during
not only utilize specific accessory pigments of the renieratene deposition of the Pesciara sediments was not enhanced and did not
carotenoid type but also employ bacteriochlorophyll for photosynthesis partake in forming the Konservat-Lagerstätte. A different picture was
that upon early diagenesis yields butyl- and isobutyl maleimides instead obtained for the Solnhofen limestones where a variety of molecular
of the ubiquitous methyl-etyhlmaleimides (Grice et al., 1996b). Whereas indicators indicated higher salinity during deposition, including presence
the latter were present in polar lipids from the Pesciara site the former, of gammacerane, organic sulfur compounds, alkylated chromans,
indicative of photic zone euxinia could not be detected. The molecular exceptionally low pristane/phytane ratios and high proportions of
signatures thus indicate that the environment of deposition at the ββ-steranes and earlier diagenetic intermediates (Schwark et al., 1998).
Pesciara Section was strongly anoxic but did not reach the euxinic stage,
except for short intervals during deposition of the L1 level. 5.5. Stratigraphic variability

5.4. Paleosalinity In order to assess the stratigraphic variability along the profile using
biomarkers it is mandatory that these compounds are preserved in situ
A variety of paleosalinity indicators based on molecular composi- and no migration or impregnation occurred. This pre-assumption is not
tion of extractable bitumens have been established (ten Haven et al., guaranteed for samples 0203 and 0204 which are prone to dissolution
1985; Peters et al., 2005). Enhanced salinity environments do support and migration of aqueous and hydrocarbon fluids. Upon migration of
highly variable microbial communities depending on the degree of aliphatic hydrocarbons fractionation occurs that will lead to a
salinity and other environmental variables. Consequently, a large preferential retardation of branched compound in the source interval
variety of molecular indicators of enhanced salinity have been and an enrichment of n-alkanes in the impregnated pore network of the
postulated (for summary see e.g. ten Haven et al., 1985; de Leeuw host rock (Leythaeuser and Schwarzkopf, 1986). Samples 0203 and 0204
and Sinnighe Damsté, 1990; Schwark et al., 1998; Peters et al., 2005). In from the Pesciara Section do exhibit impregnation features (Table 3,
many cases microbial communities that are able to cope with Fig. 5) and thus are excluded from most of the following discussion as
enhanced salinity environments also are able to survive in other the autochthony of the biomarkers is questionable.
hostile, e.g. euxinic environments. Microbial communities in many Along the profile a notable change occurs at about 7 m above datum,
cases thrive at interface layers in the water column, which can be i.e. between samples 0207 and 0212. The lower part of the section is
haloclines, thermoclines, pycnoclines or chemoclines (oxic/anoxic characterized by higher proportions of marly, laminated limestones
interfaces) or combinations thereof. Microbial communities living at lacking benthic fauna. The upper part of the section is characterized by a
such interfaces in stratified water bodies are difficult to distinguish by dominance of biocalcarenites and biocalcirudites with abundant benthic
biomarkers. As a consequence no single biomarker ratio is suitable for element (Papazzoni and Trevisani, 2006). This points towards an
distinction of enhanced salinity but a variety of parameters should be increase in basin ventilation and higher oxygen availability in the
applied to increase reliability of interpretation. upper section and more restricted and anoxic conditions in the lower
A biomarker often applied as paleosalinity indicator is the triterpe- section.
noid gammacerane (ten Haven et al., 1985; Schwark et al., 1998; The pristane/phytane ratio is a well established indicator of redox
Stephens and Carroll, 1999) that derives from tetrahymanol, a conditions (Powell and McKirdy, 1973; Didyk et al., 1978) that depends
compound biosynthesized by bacteriovorous ciliates. Gammacerane on the differentiation of phytol degradation under reducing versus
has also been reported from non-hypersaline settings and may in oxidizing conditions. Oxygen availability leads to formation of
general be an indicator for strictly stratified water bodies (Sinninghe phytenic acid which is consecutively decarboxylated thus yielding
Damsté et al., 1995), a situation often coinciding with salinity strati- pristane, an isoprenoid that has lost one carbon atom when compared
fication. Gammacerane was not detected in significant amounts in the to its biological precursor phytol. Care must be exercised in the
total extracts from the Pesciara Section. interpretation of pristane/phytane ratios, if one of the components is
Commonly very low pristane/phytane ratios of b0.5 have been not originating from chlorophyll breakdown but derived from other
associated with hypersalinity (ten Haven et al., 1985, 1987; Barbé et al., sources (ten Haven et al., 1987; Frimmel et al., 2004). In particular
1990; Schwark et al., 1998), though it is clear that this ratio is also very tocopherol derived from marine algae may increase the relative
susceptible to the redox regime. The pristane/phytane ratio in the proportion of pristane (Goossens et al., 1984; Frimmel et al., 2004),
Pesciara sediments analyzed varies between 0.55 and 1.0 and thus whereas phytane may originate from breakdown of bisphytanes (ten
indicates highly reducing conditions as stated above. In order to indicate Haven et al., 1987). No indication of additional sources for pristane and
282 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

phytane could be detected in the Pesciara Section. The ratio may thus C29 HBI (Rowland and Robson, 1990; Sinnighe Damsté et al., 2004; Peters
be taken to reflect changes in Eh condition during sedimentation. et al., 2005), the occurrence of the C29 HBI may be explained by
A marked shift occurs at 7 m above datum where pristane/phytane degradation of a C30 HBI. The coeval shifts in a variety of biomarker ratios
ratios of about 0.55 to 0.7 change to values of 0.85 to 1.0 (Fig. 5, Table 4). in a transition zone 7 m above datum (Fig. 5, Table 4) may thus reflect a
This indicates a general shift in the redox regime, switching from response to environmental change inducing establishment of a different
severely reducing conditions to still strongly reducing conditions at microbial association and/or a change in preservation conditions.
about 7 m above the base of the L1 horizon. As discussed in the section
on redox conditions, a good correlation between the pristane/phytane 6. Conclusions
and phytane/n-C18 ratio is noted for the Pesciara site, which is
confirmed by elevated phytane/n-C18 ratios in the lower part of the Investigation of kerogen and extractable lipid biomarker composi-
section (Fig. 5). tion of 13 samples taken along a stratigraphic profile of the Eocene
The change in environmental conditions is not reflected in the bulk Konservat-Lagerstätte of Pesciara di Bolca revealed new details on
parameters, which show trends mostly affected by lithology or organic matter input, its preservation conditions and factors influen-
texture. The HI values are enriched in samples 0201, 0202, 0207, and cing the preservation of delicate fossil remains. These data allow to
0218, all of which are laminated micritic limestones. It is important to confirm and refine the paleoenvironmental model presented by
note that laminitic sample 0218 is positioned in the upper section of Papazzoni and Trevisani (2006). Due to the small size and fragmenta-
the Pesciara profile and according to its biomarker signature is notably tion of the outcrops in the Bolca area, a generalized depositional
different from the underlying laminitic beds. This exemplifies that the model for the Pesciara is still affected by a speculative component.
biomarker signal is not coupled to the bulk organic matter composi- Nevertheless, accumulating evidence continues to fit the model,
tion and is suitable to delineate more delicate changes in environ- which is coherent both with the paleoecologic data from the literature
mental conditions. and with the geochemical and facies analyses.
The δ13C signature of the bulk organic matter also shows a vertical The predominant origin of kerogenous organic matter in the
trend of heavier δ13C values in the lower section and lighter δ13C Pesciara sediments can be attributed to marine organisms with a
values in the upper section (Fig. 5), superimposed on this trend is a minor admixture of terrigenous material. Biomarkers reveal that the
tendency of heavier δ13C signatures for laminated micrites and lighter terrigenic fraction is predominantly made up by land plant waxes that
δ13C values for the arenitic samples. were transported into the basin by eolian processes. The geochemical
The differentiation between upper and lower section at the data are in accordance with macrofossil remains of terrigenous plant
Pesciara site is further reflected by the amount of terrigenous material remains (Massalongo, 1859a,b), pollen, spores, and amber nodules
brought into the basin. The lower section shows a higher proportion of (Trevisani et al., 2005). A preferential enrichment of terrigenous
land plant wax derived long-chain alkanes, whereas the upper section derived organic matter is noted for the only organic-rich (8.6% TOC)
is dominated by short-chain aquatic-derived n-alkanes (Fig. 5, sample from the basal fish-bearing L1 level, indicating the importance
Table 4). This change in organic matter supply may be responsible of terrestrial freshwater run-off at this stage. Cyclic biomarkers show
for the heavier δ13C values in the lower section. Influx of terrestrial evidence for angiosperm but not for gymnosperm derived lipids.
organic matter must be accompanied by a hydrodynamic regime that Molecular biomarkers indicate that the marine organic production
allowed for transport of nutrients into the basin thus favoring higher was dominated by diatoms. The lack of micropaleontological evidence
bioproductivity during deposition of sediments in the lower section. A for diatom abundance may have arisen from dissolution of diatom tests
relationship between the autochthonous biomarker signals and the under alkaline pore water conditions; the dissolved biogenic silica could
allochthonous long-chain n-alkanes is thus established. be in turn the source of the observed silicified levels in different parts of
The biomarkers originating from aquatic organisms, i.e. algae or the Pesciara succession. Molecular biomarkers do not allow differentia-
cyanobacteria, reflect changes in environmental conditions by the tion between benthic and planktonic diatoms, thus a potential role of
variability in short-chain n-alkane composition and the ratio of n-alkanes benthic diatom mats in the preservation process of fossils cannot be
versus isoprenoids and HBI. The algal community has changed at the proven. It may be speculated that benthic diatoms in the Pesciara basin
boundary between upper and lower section in that a notable predomi- may have fulfilled the role of substrate stabilization and fossil
nance of n-C16 is established in the upper interval (Fig. 5). However, as preservation that in other settings has been ascribed to cyanobacterial
both n-C16 and n-C17 can be biosynthesized by various classes of algae mats. The latter are assumed to have been of low abundance due to the
and cyanobacteria (Peters et al., 2005), a clear source allocation based on moderate amounts of mid-chain branched alkanes and other cyano-
exclusively these two compounds is not yet possible. Compound-specific bacterial biomarkers.
isotope analyses of both n-alkanes might provide a means of separating The environment of deposition in the Pesciara basin is inferred to have
cyanobacterial from algal input. The change in aquatic community been anoxic though not euxinic, with a water column stratification that
structure is also seen in the relative abundance of n-C17 versus the diatom was due to episodic freshening of the upper water layers by monsoonal
derived HBI20, the latter being dominant in the lower section (Fig. 5). The rain and terrestrial run-off, as shown by terrigenous biomarkers and
shift in aquatic primary producers is reflected by several biomarker ratios fossils (Fig. 6). Stratification due to establishment of hypersaline bottom
indicative of the primarily source controlled nature of the environmental waters as a consequence of highly evaporitic conditions on an equatorial
change at about 7 m above datum. The laminitic micrite of sample 0214 carbonate platform, as proposed for the Solnhofen type deposits, could
provides evidence that the changes across the profile are not simply not be deduced from organic matter composition.
governed by lithofacies as this sample, despite its lithology being similar A marked shift in environmental conditions and community
to the lower section, shows a lower proportion of terrigenous long-chain structure was detected about 7 m above the datum, i.e. between
n-alkanes as well as diatom derived HBI. fossil-bearing levels L2 and L3. Diminished input of terrigenous organic
A shift in the relative proportions of HBI with 25 versus 29 carbon matter from this transition zone upwards was accompanied by less
atoms may attest to either a change in the diatom community that reducing conditions. This indicates that water column stratification
inhabited the Pesciara basin (Fig. 5) or may indicate a shift in preservation and associated anoxia in the lower part of the Pesciara Section were
conditions. As no organisms have been reported so far to biosynthesize maintained by episodic establishment of a freshwater layer on top of

Fig. 6. 4Revised depositional model for the Pesciara Section, according to previous studies (Trevisani et al., 2005; Papazzoni and Trevisani, 2006) and data from this work. A) Depositional
setting during lowstand times, with effective threshold and permanent water stratification; these conditions are recorded in the lower part of the Pesciara Section. B) Depositional setting
during highstand times, with less effective threshold and temporary water stratification; these conditions are recorded in the upper part of the Pesciara Section.
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 283
284 L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285

unmixed oxygen depleted deep waters. The topography with a Bosellini, A., 1989. Dynamics of Tethyan carbonate platform. In: Crevello, P.D., James, L.W.,
Sarg, J.F., Read, J.F. (Eds.), Controls on Carbonate Platform and Basin Platform, vol.
circulation inhibiting submarine threshold favored stratification. The 44. Society of Economic Paleontologists and Mineralogists, Special Publication,
decreasing input of terrestrial organic matter (mainly angiosperm pp. 3–13.
waxes) from 7 m upwards could be linked to a decrease in exposure of Briggs, D.E.G., 2003. The role of decay and mineralization in the preservation of soft-
bodied fossil. Annual Reviews in Earth and Planetary Sciences 31, 275–301.
emerged lands, due to a relative highstand developed after the Brocks, J.J., Summons, R.E., 2007. Sedimentary hydrocarbons, biomarkers for early life.
deposition of L2. In contrast, the lower part of the Pesciara succession In: Holland, H.D., Turekian, K.K. (Eds.), Treatise on Geochemistry, vol. 8, pp. 64–103.
could represent a relative lowstand, with increased amount of lands Buatois, L.A., Mángano, M.G., Fregenal-Martínez, M.A., de Gibert, J.M., 2000. Short-term
colonization trace-fossil assemblages in a carbonate lacustrine Konservat-Lager-
leading to greater amounts of angiosperm plants (source of the waxes
stätte (Las Hoyas Fossil Site, Lower Cretaceous, Cuenca, Central Spain). Facies 43,
preserved in the basin). 145–156.
This interpretation is confirmed by facies analysis, especially by Charbonnier, S., Vannier, J., Gaillard, C., Bourseau, J.-P., Hantzpergue, P., 2007. The La
Voulte Lagerstätte (Callovian): evidence for a deep water setting from sponge and
major thickness and continuity of L1 and L2 as compared with L3 and
crinoid communities. Palaeogeography, Palaeoclimatology, Palaeoecology 250,
L4. The relative sea-level lowstand maximized the effectiveness of the 216–236.
threshold, inducing stagnation and very restricted circulation in the Dal Degan, D., Barbieri, S., 2003. Rilievo geologico dell'area di Bolca (Monti Lessini
deeper parts of the basin (Fig. 6A). On the contrary, during relative orientali). Bollettino del Museo Civico di Storia Naturale di Verona 27, 3–10.
de Leeuw, J.W., Sinninghe Damsté, J.S., 1990. Organic sulfur compounds and other
sea-level highstand intervals the threshold was less effective, allowing biomarkers as indicators of palaeosalinity. Geochemistry of Sulfur in Petroleum
to some extent admixing and slight oxygenation of the bottom waters Systems, ACS Symposium Series 429, 417–443.
with lower Eh values compared with the former (Fig. 6B). De Zanche, V., Conterno, T., 1972. Contributo alla conoscenza geologica dell'orizzonte di
Roncà nel Veronese e nel Vicentino. Atti e Memorie dell'Accademia Patavina di
Since the whole Pesciara succession has been dated to the upper Scienze, Lettere ed Arti 84, 287–295.
part of the SBZ 11 (Papazzoni and Trevisani, 2006), and the overall Didyk, B.M., Simoneit, B.R.T., Brassell, S.C., Eglinton, G., 1978. Organic geochemical
duration of this biozone is 1.1–1.2 Ma (Serra-Kiel et al., 1998), we can indicators of palaeoenvironmental conditions of sedimentation. Nature 272, 216–222.
Espitalié, J., Laporte, J.L., Madec, M., Marquis, F., Leplat, P., Paulet, J., Boutefeu, A., 1977.
interpret the body of the Pesciara as a parasequence of the 4th order Méthode rapide de caractérisation des roches mères de leur potentiel pétrolier et
(0.01–0.5 Ma). de leur degré d'évolution. Revue de l'Institut Francais du Pétrole 32, 23–42.
In turn, this parasequence is subdivided into at least four higher- Frimmel, A., Oschmann, W., Schwark, L., 2004. Chemostratigraphy of the Posidonia
Black Shale, SW-Germany: I — Influence of sea level variation on organic facies
order parasequences generated by minor sea-level fluctuations. Each evolution. Chemical Geology 206, 199–230.
of these higher-order parasequences is represented by a couple of Fürsich, F.T., Mäuser, M., Schneider, S., Werner, W., 2007. The Wattendorf Plattenkalk
laminated micrite with fish (lowstand deposits, LD) — biocalcarenites (Upper Kimmeridgian) — a new Conservation Lagerstätte from the northern
Franconian Alb, southern Germany. Neues Jahrbuch Geologie und Paläontologie
with abundant benthic fossils (highstand deposits, HD). The trans-
245, 45–58.
gressive deposits (TD) are clearly represented at the top of L1 by an Gagosian, R.B., Peltzer, E.T., 1986. The importance of atmospheric input of terrestrial
undulated bed, 0 to 30-cm thick, of bioclastic limestone with silicified organic material to deep-sea sediments. Organic Geochemistry 10, 661–669.
extraclasts with alveolinas of the SBZ 10. Gagosian, R.B., Peltzer, E.T., Merrill, J.T., 1987. Long-range transport of terrestrially
derived lipids in aerosols from the south Pacific. Nature 325, 800–804.
Goossens, H., de Leeuw, J.W., Schenck, P.A., Brassell, S.C., 1984. Tocopherols as likely
Acknowledgements precursors of pristane in ancient sediments and crude oils. Nature 312, 440–442.
Grice, K., Schaeffer, P., Powell, S., Schwark, L., Maxwell, J., 1996a. Molecular indicators of
paleoenvironmental conditions in an immature Permian Shale (Kupferschiefer,
We thank Dr L. Langone (CNR-ISMAR Bologna), Dr D. Malferrari Lower Rhine Basin, Germany) from free and sulfide-bound lipids. Organic
(Modena University) and Dr M. Tonelli (CIGS; Modena University) for Geochemistry 25, 131–147.
technical support during this study. Prof. A. Rossi (Modena University) Grice, K., Gibbison, R., Atkinson, J.E., Schwark, L., Eckardt, C.B.E., Maxwell, J.R., 1996b.
Maleimides (1H-pyrrole-2,5-diones) as indicators of anoxygenic photosynthesis in
provided X-ray analyses and helpful comments on the results. ancient water columns. Geochimica et Cosmochimica Acta 60, 3913–3924.
Excellent analytical work by student technical assistants at Cologne Hefter, J., Thiel, V., Jenisch, A., Galling, U., Kempe, S., Michaelis, W., 1993. Biomarker
University is gratefully acknowledged. indications for microbial contribution to Recent and Late Jurassic carbonate
deposits. Facies 29, 93–106.
This study was supported by funds from the Ministero dell'Uni- Hückel, U., 1974. Vergleich des Mineralbestandes der Plattenkalke Solnhofens und des
versità e della Ricerca Scientifica (Resp. A. Ferretti), by Società Libanon mit anderen Kalken. Neues Jahrbuch für Geologie und Paläontologie
Paleontologica Italiana grant (Fondo per la Tutela del Patrimonio Abhandlungen 145, 153–182.
Keupp, H., Jenisch, A., Herrmann, R., Neuweiler, F., Reitner, J., 1993. Microbial carbonate
Paleontologico), and by Museo Civico di Storia Naturale di Ferrara.
crusts — a key to the environmental analysis of fossil spongiolites? Facies 29, 41–54.
Killops, S., Killops, V., 2005. Introduction to Organic Geochemistry, 2nd edition.
Blackwell Publishing. 393 pp.
References Langford, F.F., Blanc-Valleron, M.M., 1990. Interpreting Rock-Eval pyrolysis data using
graphs of pyrolysable hydrocarbons vs. total organic carbon. AAPG Bulletin 74,
Allison, P.A., 1988. Konservat-Lagerstätten: cause and classification. Paleobiology 14, 799–804.
331–344. Leythaeuser, D., Schwarzkopf, T., 1986. The pristane/n-heptadecane ratio as an indicator
Allison, P.A., Briggs, D.E.G., 1993. Exceptional fossil record: distribution of soft-tissue for recognition of hydrocarbon migration effects. Organic Geochemistry 10,191–197.
preservation through the Phanerozoic. Geology 21, 527–530. Li, M., Larter, S.R., Taylor, P., Jones, D.M., Bowler, B., Bjoroy, M., 1995. Biomarkers or not
Barbé, A., Grimalt, J.O., Pueyo, J.J., Albaiges, J., 1990. Characterization of model evaporitic biomarkers? A new hypothesis for the origin of pristane involving derivation from
environments through study of lipid components. Organic Geochemistry 16, 815–828. methyltrimethyltridecylchromans (MTTCs) formed during diagenesis from chlor-
Barbieri, G., 1972. Sul significato geologico della Faglia di Castelvero (Lessini veronesi). ophyll and alkylphenols. Organic Geochemistry 23, 159–167.
Atti e Memorie dell'Accademia Patavina di Scienze, Lettere ed Arti 84, 297–302. Lüniger, G., Schwark, L., 2002. Organofacies and early diagenesis of organic matter from
Barbieri, G., Medizza, F., 1969. Contributo alla conoscenza geologica della regione di the Oligocene lacustrine Enspel Oil Shale Deposit, Germany. Journal of Sedimentol-
Bolca (Monti Lessini). Memorie dell'Istituto di Geologia e Mineralogia dell'Uni- ogy 148, 275–288.
versità di Padova 27, 1–36. Marschner, B., Brodowski, S., Dreves, A., Gleixner, G., Grootes, P.M., Hamer, U., Heim, A.,
Barbieri, G., De Zanche, V., Medizza, F., Sedea, R., 1982. Considerazioni sul vulcanismo Jandl, G., Ji, R., Kaiser, K., Kalbitz, K., Kramer, C., Rethemeyer, J., Schmidt, M.W.I.,
terziario del Veneto occidentale e del Trentino meridionale. Rendiconti della Schwark, L., Wiesenberg, G.L.B., 2008. Is recalcitrance a relevant property for the
Società Geologica Italiana 4 (1981), 267–270. stabilization of organic matter in soils? Journal of Plant Nutrition and Soil Science
Barbieri, G., De Zanche, V., Sedea, R., 1991. Vulcanismo paleogenico ed evoluzione del 171, 1–20.
semigraben Alpone-Agno (Monti Lessini). Rendiconti della Società Geologica Massalongo, A., 1859a. Specimen photographicum quorumdam plantarumque
Italiana 14, 5–12. fossilium agri veronensis. Vicentini-Franchini, Verona.
Barbieri, G., Zampieri, D., 1992. Deformazioni sinsedimentarie eoceniche con stile a Massalongo, A., 1859b. Syllabus plantarum fossilium hucusque in formationibus
domino nel semigraben Alpone-Agno e relativo campo di paleostress (Monti Lessini tertiariis Agri Veneti detectarum. Merlo, Verona.
orientali-Prealpi Venete). Atti Ticinensi di Scienze della Terra 35, 25–31. Medizza, F., 1975. Il nannoplancton calcareo della Pesciara di Bolca (Monti Lessini). Studi
Barthel, K.W., Swinburne, N.H.M., Conway Morris, S., 1990. Solnhofen — a Study in e Ricerche sui Giacimenti Terziari di Bolca 2, 433–444.
Mesozoic Palaeontology. Cambridge University Press. 236 pp. Papazzoni, C.A., Ferretti, A., Trevisani, E., 2007. The Eocene Fossil-Lagerstätte of Bolca
Bordenave, M.L., Espitalié, J., Leplat, P., Oudin, J.L., Vandenbroucke, M., 1993. Screening (Italy): a guarantee of organic matter preservation? European Geosciences Union
techniques for source rock evaluation. In: Bordenave, M.L. (Ed.), Applied Petroleum General Assembly. Vienna, Austria, 15–20 April 2007, Geophysical Research
Geochemistry. Editions Technip, Paris, pp. 219–278. Abstracts 9, 08157. 2 pp.
L. Schwark et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 272–285 285

Papazzoni, C.A., Trevisani, E., 2006. Facies analysis, palaeoenvironmental reconstruc- Sinninghe Damsté, J.S., Kohnen, M.E.L., de Leeuw, J.W., 1990. Thiophenic biomarkers for
tion, and biostratigraphy, of the “Pesciara di Bolca” (Verona, northern Italy): an palaeoenvironmental assessment and molecular stratigraphy. Nature 345, 609–611.
early Eocene Fossil-Lagerstätte. Palaeogeography, Palaeoclimatology, Palaeoecology, Sinninghe Damsté, J.S., Keely, B.J., Betts, S.E., Baas, M., Maxwell, J.R., de Leeuw, J.W., 1993.
242 (1–2), 21–35. Variations in abundances and distributions of isoprenoid chromans and long-chain
Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed alkylbenzenes in sediments of the Mulhouse Basin: a molecular sedimentary record
pyrolysis. AAPG Bulletin 70, 318–329. of palaeosalinity. Organic Geochemistry 20, 1201–1215.
Peters, K.E., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide, Second ed. Sinninghe Damsté, J.S., Kenig, F., Koopmans, M.P., Köster, J., Schouten, S., Hayes, J.M., de
Cambridge University Press, Cambridge. Leeuw, J.W., 1995. Evidence for gammacerane as an indicator of water column
Powell, T.G., McKirdy, D.M., 1973. Relationship between the ratio of pristane to phytane, stratification. Geochimica et Cosmochimica Acta 59, 1895–1900.
crude-oil composition and geological environment in Australia. Nature 243, 37–39. Sinninghe Damsté, J.S., Muyzer, G., Abbas, B., Rampen, S.W., Masse, G., Allard, W.G., Belt,
Rontani, J.F., Volkman, J.K., 2005. Lipid characterization of coastal hypersaline S.T., Robert, J.M., Rowland, S.J., Moldowan, J.M., Barbanti, S.M., Fago, F.J., Denisevich,
cyanobacterial mats from the Camargue (France). Organic Geochemistry 36, P., Dahl, J., Trindade, L.A.F., Schouten, S., 2004. The rise of the rhizosolenid diatoms.
251–272. Science 304, 584–587.
Rowland, S.J., Robson, J.N., 1990. The widespread occurrence of highly branched acyclic Sorbini, L., 1972. I fossili di Bolca. Museo Civico di Storia Naturale di Verona, Verona.
C-20, C-25 and C-30 hydrocarbons in recent sediments and biota — a review. 133pp.
Marine and Environmental Research 30, 191–216. Stephens, N.P., Carroll, A.R., 1999. Salinity stratification in the Permian Phosphoria sea; a
Schwark, L., Frimmel, A., 2004. Chemostratigraphy of the Posidonia Black Shale, SW- proposed paleoceanographic model. Geology 27, 899–902.
Germany: II — Assessment of extent and persistence of Photic Zone Anoxia via aryl Summons, R.E., Powell, T.G., 1986. Chlorobiaceae in Palaeozoic seas revealed by
isoprenoid distributions. Chemical Geology 206, 231–248. biological markers, isotopes and geology. Nature 319, 763–765.
Schwark, L., Püttmann, W., 1990. Aromatic hydrocarbon composition of the Permian Swinburne, N.H.M., Hemleben, C., 1994. The plattenkalk facies: a deposit of several
Kupferschiefer in the Lower Rhine Basin, NW Germany. Organic Geochemistry 16, environments. Geobios 27, 313–320.
749–761. ten Haven, H.L., de Leeuw, J.W., Schenk, P.A., 1985. Organic geochemical studies of a
Schwark, L., Vliex, M., Schaeffer, P., 1998. Geochemical characterization of Malm Zeta Messinian evaporitic basin, northern Apennines (Italy). I: Hydrocarbon biomarkers
laminated carbonates from the Franconian Alb, SW-Germany (II). Organic for a hypersaline environment. Geochimica et Cosmochimica Acta 49, 2181–2191.
Geochemistry 29, 1921–1952. ten Haven, H.L., de Leeuw, J.W., Rullkötter, J., Sinninghe Damsté, J.S., 1987. Restricted
Seilacher, A., 1970. Begriff und Bedeutung der Fossil-Lagerstätten. Neues Jahrbuch utility of the pristane/phytane ratio as a palaeoenvironmental indicator. Nature
Geologie und Paläontologie Abhandlungen, 34–39. 330, 641–643.
Seilacher, A., 1990. Taphonomy of Fossil-Lagerstätten: an overview. In: Briggs, D.E.G., Trevisani, E., Papazzoni, C.A., Ragazzi, E., Roghi, G., 2005. Early Eocene amber from the
Crowther, P. (Eds.), Paleobiology. Blackwell Science, pp. 266–270. “Pesciara di Bolca” (Lessini Mountains, Northern Italy). Palaeogeography, Palaeo-
Serra-Kiel, J., Hottinger, L., Caus, E., Drobne, K., Ferràndez, C., Jauhri, A.K., Less, G., climatology, Palaeoecology 223 (3–4), 260–274.
Pavlovec, R., Pignatti, J., Samsó, J.M., Schaub, H., Sirel, E., Strougo, A., Tambareau, Y., Tyson, R.V., 1995. Sedimentary Organic Matter: Organic Facies and Palynofacies.
Tosquella, J., Zakrevskaya, E., 1998. Larger foraminiferal biostratigraphy of the Chapman and Hall, London. 615pp.
Tethyan Paleocene and Eocene. Bulletin de la Société Géologique de France 169, Yon, D.A., Maxwell, J.R., Ryback, G., 1982. 2,6,10-Trimethyl-7-(3-methylbutyl)-dodecane,
281–299. a novel sedimentary biological marker compound. Tetrahedron Letters 23,
Sinninghe Damsté, J.S., Kock-van Dalen, A.C., de Leeuw, J.W., Schenck, P.A., Guoying, S., 2143–2146.
Brassell, S.C., 1987. The identification of mono-, di- and trimethyl 2-methyl-2- Zampieri, D., 1992. Extensional mesostructures in the Cretaceous limestones of
(4,8,12-trimethyltridecyl)chromans and their occurrence in the geosphere. Geochi- Paleogene Alpone-Agno half-graben (Eastern Lessini Mountains, Northern Italy).
mica et Cosmochimica Acta 51, 2393–2400. Studi Trentini di Scienze Naturali, Acta Geologica 67 (1990), 75–85.
Sinninghe Damsté, J.S., Rijpstra, I., de Leeuw, J.W., Schenck, P.A., 1989. The occurrence Zampieri, D., 1995. Tertiary extension in the southern Trento Platform, Southern Alps,
and identification of series of organic sulfur compounds in oils and sediment Italy. Tectonics 14, 645–657.
extracts. II. Their presence in samples from hypersaline and non-hypersaline
palaeoenvironmental and maturity indicators. Geochimica et Cosmochimica Acta
53, 1323–1341.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Geochemical environment of the Coniacian–Santonian western tropical Atlantic at


Demerara Rise
C. März a,⁎, B. Beckmann b,1, C. Franke c, Christoph Vogt a, T. Wagner d, S. Kasten e
a
Department of Geosciences, University of Bremen, Klagenfurter Str., 28359 Bremen, Germany
b
Institute for Geology and Mineralogy, University of Cologne, Zülpicher Str. 49a, 50674 Cologne, Germany
c
Laboratoire des Sciences du Climat et de l'Environnement (LSCE), CEA-CNRS-UVSQ, 12 Avenue de la Terrasse, 91198 Gif-sur-Yvette, France
d
School of Civil Engineering and Geosciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, UK
e
Alfred Wegener Institute for Polar and Marine Research, Am Handelshafen 12, 27570 Bremerhaven, Germany

A R T I C L E I N F O A B S T R A C T

Article history: Organic carbon-rich shales deposited during the Coniacian–Santonian Oceanic Anoxic Event 3 were drilled
Received 13 December 2007 during ODP Leg 207 at Demerara Rise. We present integrated high-resolution geochemical records of core
Received in revised form 13 May 2008 intervals from ODP Sites 1259 and 1261 both from nannofossil biozone CC14. Our results reveal systematic
Accepted 16 May 2008
variations in marine and detrital sediment contribution, depositional processes, and bottom water redox
conditions during black shale formation at two locations on Demerara Rise in different paleo-water depths. A
Keywords:
Oceanic anoxic event
combination of redox proxies (Fe/S, P/Al, C/P, redox-sensitive/sulfide-forming trace metals Mn, Cd, Mo, Ni, V,
Trace elements Zn) and other analytical approaches (bulk sediment composition, P speciation, electron microscopy, X-ray
Iron diffraction) evidence anoxic to sulfidic bottom water and sediment conditions throughout the deposition of
Phosphorus black shale. These extreme redox conditions persisted and were periodically punctuated by short-termed
Sediment source periods with less reducing bottom waters irrespective of paleo-water depth. Sediment supply at both sites
Redox changes was generally dominated by marine material (carbonate, organic matter, opal) although relationships of
Enrichment factors detrital proxies as well as glauconitic horizons support some influence of turbidites, winnowing bottom
currents and/or variable detritus sources, along with less reducing bottom water at the proposed shallower
location (ODP Site 1259). At Site 1261, located at greater paleo-depth, redox fluctuations were more regular,
and steady hemipelagic sedimentation sustained the development of mostly undisturbed lamination in the
sedimentary record. Strong similarities of the studied deposits exist with the stratigraphic older
Cenomanian–Turonian OAE2 black shale sections at Demerara Rise, suggesting that the primary mechanisms
controlling continental supply and ocean redox state were time-invariant and kept the western equatorial
Atlantic margin widely anoxic over millions of years.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Danelian, 2006; Forster et al., 2007a; Junium and Arthur, 2007;
Nederbragt et al., 2007) whereas far less is known about the
Since Schlanger and Jenkyns (1976) introduced the original development of the later Coniacian–Santonian OAE3 in this critical
concept of Oceanic Anoxic Events (OAEs), much progress has been area of the tropical North Atlantic (Friedrich and Erbacher, 2006;
made in the understanding of the widespread deposition of organic Beckmann et al., in press; März et al., in press). Given that OAE3
matter (OM)-rich sediments in the Cretaceous Proto-Atlantic and sediments from the adjacent tropical African margin (Deep Ivory
adjacent seas. During ODP Leg 207, up to 90 m thick Cretaceous black Basin, ODP Leg 159) revealed spectacular evidence of astronomically
shale deposits, including OAE2 and OAE3 were recovered from forced sedimentary and redox cycles directly associated with African
Demerara Rise (Erbacher et al., 2004; Mosher et al., 2007). climate variability (Wagner, 2002; Wagner et al., 2004; Beckmann
Cenomanian–Turonian OAE2 black shales from ODP Leg 207 have et al., 2005a,b; Flögel and Wagner, 2006) and the opening of the
been subject of intensive research (e.g. Erbacher et al., 2005; Friedrich Equatorial Atlantic (Wagner and Pletsch, 1999; Pletsch et al., 2001)
et al., 2006; Hardas and Mutterlose, 2006; Musavu-Moussavou and there is scope for more detailed investigation of this stratigraphic
interval at Demerara Rise. First geochemical studies revealed that
Demerara Rise black shales contain thermally immature marine OM
⁎ Corresponding author. Present address: ICBM, University of Oldenburg, Carl-von- (Meyers et al., 2006; Forster et al., 2007b). Its slow degradation still
Ossietzky-Str. 9-11, 26111 Oldenburg, Germany. Tel.: +49 441 798 3627.
produces methane, which is anaerobically oxidized at the top of the
E-mail address: cmaerz@icbm.de (C. März).
1
Present address: Federal Institute for Geosciences and Natural Resources, Stilleweg Cretaceous black shale succession via sulfate reduction, resulting in
2, 30655 Hannover, Germany. sulfate depletion and long-lasting barite remobilization within the

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.05.004
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 287

black shale succession (Erbacher et al., 2004; Arndt et al., 2006). High here, however, we emphasize that this study takes the discussion
degrees of pyritization, but also sulfurized OM (Böttcher et al., 2006), forward by comparing synchronous stratigraphic intervals from two
enrichments of Mo, V, Zn, and Mn depletion (Hetzel et al., 2006) different paleo-water depths, and by adding new geochemical
indicate overall oxygen-depleted, periodically even sulfidic bottom information on sediment sources and depositional processes. To
waters. Enrichments of P as well as high amounts of Ba (despite place the results from OAE3 into wider perspective we compare them
diagenetic redistribution) support enhanced nutrient supply and high with new geochemical data from OAE2 at Demerara Rise (Brumsack,
productivity during black shale deposition (Arndt et al., 2006; Hetzel 2006; Junium and Arthur, 2007; Hetzel et al., 2009).
et al., 2006). We recently explored a broad variety of redox proxies in
OAE3 black shales at ODP Site 1261. The records are consistent and 2. Material and methods
show cyclic and rapid redox fluctuations from sulfidic to anoxic, non-
sulfidic conditions in bottom waters (März et al., in press) obviously The ODP Leg 207 Sites 1259 and 1261 were drilled on Demerara
triggered astronomically by short eccentricity cycles (Flögel et al., in Rise, in water depths of 2354 m and 1899 m on the northern and
press). northwestern slope, respectively (Fig. 1; Erbacher et al., 2004). For this
In this study we add evidence to the findings of the Cretaceous low high-resolution study, we chose core intervals from Sites 1259
latitude Atlantic by applying a wide analytical spectrum in high (499.60–500.61 meters composite depth, mcd, 101 cm) and 1261
temporal resolution to ODP Sites 1259 and 1261 from Demerara Rise. (570.20–571.40 mcd, 120 cm) and sampled them in continuous 1 cm
The main goal is to trace changes in deep ocean redox state, sediment resolution. For comparison we selected intervals from the same
sources and depositional processes along a bathymetric transect stratigraphic nannofossil biozone CC14 (Coniacian–Santonian), using
across the S-American continental margin. To achieve these goals we the nannofossil stratigraphy published by Erbacher et al. (2004) that
selected nannofossil biozone CC14 intervals from both sites and was recently refined by Flögel et al. (in press) for Site 1261 and by
applied major and trace element geochemistry, P speciation, X-ray Bornemann et al. (2008) for Site 1259. Sedimentation rates were
diffraction, and scanning electron microscopy. For completeness, ~5.4 mm/ka at Site 1261 (Flögel et al., in press), and ~2–3 mm/ka at
some of the results presented by März et al. (in press) are shown Site 1259 (Bornemann et al., 2008). Thus, both sampled intervals are

Fig. 1. a) Global plate-tectonic reconstruction for 80 Ma BP (after Hay et al., 1999; Wagner et al., 2004). Grey areas represent regions above sea level, white areas represent oceans,
black lines depict plate boundaries. Black dots indicate locations of reported Late Cretaceous organic matter-rich deposits (Wagner et al., 2004 and references therein). b) Close-up of
the equatorial Proto-Atlantic, with Demerara Rise marked by the black rectangle. c) Present-day bathymetry of Demerara Rise, with drill sites 1257–1261.
288 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

condensed successions, with sedimentation rates at Site 1261 about X-ray diffraction (XRD) was used to reveal the general mineralogy
twice as high as at Site 1259. Being aware of these differences, our of the core sections, and to support P extraction data by detecting the
geochemical comparison does not aim at layer-to-layer correlation — apatite fraction of the bulk P pool. X-ray diffractograms of freeze-dried,
although both intervals were deposited during nannofossil biozone ground samples were produced with a Philips X'Pert Pro multipurpose
CC14, there might still be a temporal offset. We rather compare both diffractometer equipped with a Cu-tube (k" 1.541, 45 kV, 40 mA), a
sites in terms of general sedimentation patterns and geochemical fixed divergence slit of 1/4°, a 16 samples changer, a secondary
characteristics of sediments and bottom waters. monochromator and the X'Celerator detector system. Samples were
Sample splits taken at the ODP Core Repository Bremen were frozen, scanned continuously from 3–85° 2θ, with a calculated step size of
freeze-dried and ground in an agate mortar. Total organic carbon (TOC) 0.016° 2θ (calculated time per step 100 s). Mineral identification and
and carbonate contents were determined with a Leco CS 200. First, total semi-quantification was done with the Philips software X'Pert High-
carbon (TC) was measured, then a sample aliquot was decalcified with Score™ (as Relative Intensity Ratio values, calculated as ratio of the
12.5% HCl, washed twice and dried at 60 °C to determine TOC. The CaCO3 most intense reflex of a specific mineral phase to most intense reflex of
content (%) was calculated as (TC − TOC) ⁎ 8.33. Calcite was the only pure corundum; Chung, 1974; Vogt et al., 2002).
carbonate mineral detected by X-ray diffraction (XRD), other carbonates Heavy liquid separation (Franke et al., 2007) was applied to
(e.g. dolomite, rhodochrosite) are not of quantitative importance selected samples, extracting the mineral fraction with a density
(b1 wt.%). Minor amounts might be present especially at Site 1259, of N3.0 g/cm3 (e.g. barite, pyrite, calcium-fluorapatite = CFA). We used
but for simplicity are summarized as CaCO3. Several standards with C ~0.5–1 g of freeze-dried, homogenized sediment and a Na-polytung-
contents between 0.5 and 12% were applied to keep the accuracy of the state solution (3.0 g/cm3; ~25 ml per sample). Sediment solutions
measurement at N97%. were shaken manually and dispersed in an ultrasonic bath. Suspen-
About 50 mg of each sample were subjected to a microwave total sions were centrifuged (20 min, 4000 rps), and the heavy liquid mirror
acid digestion procedure (3 ml HNO3, 2 ml HCl and 2 ml HF; heated up and supernatant fraction (b3.0 g/cm3) were removed with a pipette.
to ~ 210 °C in closed pressure-tight teflon containers; acids were fumed The remaining heavy fraction was vacuum-filtered (0.2 µm), and
off with a special vacuum device, dry samples were redissolved with washed several times with bi-distilled water. Filtered extracts were
HNO3). Major and minor elements were measured with an Inductively dispersed in bi-distilled water and centrifuged several times to
Coupled Plasma Optical Emission Spectrometer (ICP-OES, Perkin Elmer prevent Na-polytungstate contamination. Dried extracts were pre-
3300RL) in triplicate to assure a standard deviation of b5%. Accuracy of pared for Scanning Electron Microscopy (SEM, Philips XL30 SFEG,
the digestion and measurements was assured by running blanks and 12 kV acceleration voltage). Unconsolidated grains were dispersed
the USGS standard MAG-1 and in-house standards MAX-1 and CAMAX onto a carbon sticker on a standard Al stub, and covered with a few nm
with the sample sets. Elemental concentrations were within the 5% thick carbon layer to prevent surface charging during SEM analysis.
range of certified values for the standard materials, respective standard Secondary electron (SE) and backscattered electron (BSE) detectors
deviations of triplicate measurements were b3% for major and b5% for were used for imaging. For elemental composition, energy-dispersive
trace elements. For details, we refer to Schulz (2004) and www. X-ray spectroscopy (EDS; e.g. Goldstein et al., 1992) was applied and
geochemie.uni-bremen.de/koelling/MGCmain.html. Bulk elemental semi-quantification was achieved using the EDAX PhiRhoZ software.
contents and TOC were normalized to Al to compensate for variable
dilution of the geochemical record by biogenic and diagenetic 3. Results
components. Element contents calculated on carbonate-free basis
(CFB; data not shown), an alternative method to compensate for 3.1. Lithology
variable biogenic calcite dilution, showed very similar patterns.
A sequential phosphate extraction (SEDEX method by Ruttenberg, OAE3 sediments at Site 1261 consist of finely laminated black
1992; modified after Schenau and De Lange, 2000; Schenau et al., claystone and nannofossil claystone (Erbacher et al., 2004). Visual
2000) was applied to selected P-rich samples (500.13–500.53 mcd at inspection of the Site 1261 core section revealed black intervals of few
Site 1259 and 570.5–570.64 mcd at Site 1261), resulting in differentia- centimeters thickness containing light, roundish components of up to
tion between various phosphorus-bearing phases in the sediment a few mm in diameter. The components are elongate, lense-shaped,
(Table 1). Extracted P was analysed photometrically (Steps I, III–V) or and embedded into the stratification and bear no signs of physical
via ICP-AES (Step II). disturbance throughout the entire study interval. Similarly, at Site
1259 the studied core interval displays lamination in the upper
Table 1
~ 70 cm of the interval (499.60–500.30 mcd), as well as cyclic
Different steps of the applied sequential phosphate extraction, the respective extraction
solutions and extracted fractions of phosphate occurrence of layers with round to elliptic, whitish components of
up to a few mm in size. However, in the lower part a glauconitic
Step Reagents P component extracted
horizon (Erbacher et al., 2004; here referred to as GH) is visible as a
I 25 ml 2 M NH4Cl (pH 7, shake for 4 h), Pvar: pore water phosphate, amorphous sandy, light horizon of ~ 20 cm thickness, comprising a relatively
repetition up to 20× Ca-rich apatite precursor mineral,
phosphate loosely sorbed to carbonates
unsorted framework of white, mm- to cm-sized grains. Under the light
and clay minerals, fish bones microscope, glauconite grains from the GH appear as green to blue-
(= biogenic hydroxyapatite) green crystals, fragments or as green-whitish aggregates.
II 25 ml Na-dithionite solution, citrate- Piron: phosphate bound to iron
buffered (pH 7.5, shake for 8 h); wash (oxyhydr)oxides, including secondarily
3.2. Total organic carbon and carbonate
with 25 ml 2 M NH4Cl (shake for 2 h); oxidized Fe(II)-phophates as vivianite
wash with 25 ml dem. water (shake for
2 h) The TOC content at Site 1261 is generally higher compared to Site
III 25 ml 1 M Na-acetate solution (pH 4, Pauth: authigenic apatite (CFA, 1259 (average of 7.7 wt.% versus 6.1 wt.%, respectively; Fig. 2a), as is
for 6 h); wash with 25 ml 2 M NH4Cl francolite) the TOC/Al ratio (average of 3.6 versus 2.8; Fig. 2b) and TOC (CFB)
(shake for 2 h); wash with 25 ml dem.
water (shake for 2 h)
(average of 18.7 wt.% versus 15.8 wt.%; data not shown. The amplitude
IV 25 ml 1 M HCl (shake for 24 h); wash Pdet: detrital apatite between TOC maxima and minima is on the order of 7 wt.% at both
with 25 ml dem. water (shake for 2 h) sites, with values of 4.4–11.7 wt.% at Site 1261 and 2.9–10.2 wt.% at Site
after ignition at 550 °C 1259 (Fig. 2a), without clear repetitive patterns. Different from that,
V 25 ml 1 M HCl (shake for 24 h) Porg: phosphate bound to organic
the TOC/Al records show regular cyclic patterns at both sites, with the
matter
exception of much lower TOC/Al ratios in the lower part of the 1259
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 289

Fig. 2. Chemical composition of ODP Sites 1261, 570.2–571.4 (meters composite depth, mcd), and 1259, 499.6–500.7 mcd. (a) Total organic carbon (TOC, wt.%), (b) TOC/Al (%/%),
(c) carbonate (wt.%), (d) Al (g/kg) versus core depth. The grey-shaded areas indicate low carbonate contents, the glauconitic horizon (GH) is marked by the dotted area.

section (below 500.30 mcd; Fig. 2b). The carbonate record at Site 1261 At Site 1259, the correlation of Al with Ti is strong (R2 = 0.98), but the
displays background values of ~50–70 wt.%, interrupted by four one of Al with Zr is weaker (Fig. 4b, c; R2 = 0.63). Typical elements of
intervals of 5–10 cm thickness, each with carbonate contents of only light silicates are K and Mg (Fig. 3c, d), but they are also present in sea
20–30 wt.% (Fig. 2c), coincident with lowest TOC/Al values. In contrast, water, and Mg may be incorporated in carbonates (although no
the carbonate content at Site 1259 does not show the same systematic dolomite was detected via XRD). Comparable to Ti and Zr, relation-
variations, but varies in a more irregular manner between 30–70 wt.% ships of Al with K and Mg at Site 1261 are strong (R2 = 0.98 for Al with
(Fig. 2c). Only in the lowest part of the interval (below 500.48 mcd), K, 0.94 for Al with Mg; Fig. 4d, e). At Site 1259, correlations of Al with K
carbonate contents reach 70–100 wt.%. and Mg are overall weaker, with R2 = 0.65 for Al with K, and R2 = 0.36
for Al with Mg (Fig. 4d, e).
3.3. Detrital elements (Al, K, Mg, Ti, Zr)
3.4. Redox-sensitive trace elements (Cd, Mn, Mo, Ni, V, Zn)
Aluminum is mainly incorporated into detrital silicates in
continental margin settings like Demerara Rise and thus represents The investigated trace elements participate strongly in redox
continental input. Other typically detrital elements are K, Mg, Ti and Zr reactions and may become authigenically enriched or depleted in
(Fig. 3). At Site 1261, background Al values of 15–25 g/kg are sediments deposited under oxygen-free conditions, making them
interrupted by sharp peaks of 35–40 g/kg (Fig. 2d). Calculation on frequently applied tracers of paleoredox conditions (e.g. François,
CFB, however, removes these peaks, and documents minor fluctua- 1988; Calvert and Pedersen, 1993; Dean et al., 1997; Morford and
tions in Al (CFB; max. value 63.7 g/kg, min. value 45.4 g/kg, mean value Emerson, 1999; recent reviews by Brumsack, 2006; Tribovillard et al.,
52.8 g/kg; data not shown). A clear anti-correlation of Al with 2006). The degree of enrichment/depletion is calculated as enrich-
carbonate (R2 = 0.91; correlation coefficients always given as linear ment factor (EF) relative to a standard material (average shale = AS;
correlations; Figs. 2d, 4a) documents mutual dilution of detrital and Turekian and Wedepohl, 1961):
biogenic material. At Site 1259, Al contents are within a similar range
(10–35 g/kg; Fig. 2d), but the Al-carbonate relationship is weaker EFelement ¼ ðelement=AlÞsample =ðelement=AlÞaverage shale
(R2 = 0.59; Fig. 4a). Terrigenous elements bound to the heavy mineral
fraction include Ti and Zr (Fig. 3a, b). At Site 1261, correlation The use of AS as reference material may get stressed by non-uniform
coefficients are R2 = 0.99 for Al with Ti, and R2 = 0.93 with Zr (Fig. 4b, c). lithologies in the detrital source area (Van der Weijden, 2002), but this
290 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

Fig. 3. Chemical composition of ODP Sites 1261, 570.2–571.4 mcd, and 1259, 499.6–500.7 mcd. (a) Ti/Al, (b) Zr/Al (both as ppm/%), (c) K/Al, (d) Mg/Al (both as %/%) versus core depth
(meters composite depth, mcd). The grey-shaded areas indicate less reducing conditions, the glauconitic horizon (GH) is marked by the dotted area.

potential bias is uncritical for the presented trace metals, with low Maximum values of Cd/Al, V/Al and Zn/Al ratios are similar at both
contents in AS and/or strong incorporation into biological or sites, but Mo/Al and Ni/Al maxima are much lower at Site 1259 (Fig. 5).
authigenic processes (Brumsack, 2006). In the studied core sections As further differences between trace element/Al profiles at Site 1261,
we identified two categories of redox-sensitive trace elements. all trace element/Al ratios are lowest within the carbonate-poor layers
(grey bars), but the positions of the maxima vary. While Cd/Al, V/Al
3.4.1. Manganese and Zn/Al are highest right above the carbonate-poor layers and
Values of Mn/Al are ~ 20–45 ppm/% across the Site 1261 and the gradually decrease upsection, Mo/Al and Ni/Al continuously increase
upper Site 1259 interval (above 500.10 mcd). Values of ~ 60–90 ppm/% upcore and reach maximum values right below the next carbonate-
are reached in the lowest part of the 1259 section (below 500.48 mcd; poor layer. Thus, Mo/Al and Ni/Al exhibit rather strong correlations
Fig. 5a). The Mn/Al record displays a rather strong correlation with with the TOC/Al and S/Al records (R2 = 0.73 to 0.89; Fig. 6c–f), while
carbonate (Figs. 5a, 6a). At Site 1261, correlation coefficients are there are no significant correlations of Cd/Al, V/Al and Zn/Al with TOC/
similar for carbonate-poorer (R2 = 0.78, n = 19) and carbonate-richer Al and S/Al (R2 b 0.01). Site 1259 exhibits the same general trace
(R2 = 0.79, n = 96) samples if the carbonate content threshold is set to element/Al patterns, although trace element/Al minima (grey bars) do
55 wt.%. Respective correlations both follow positive trends, but with not match as precisely with carbonate minima as at Site 1261. Average
different slopes. At Site 1259, the Mn/Al to carbonate correlation is trace element EFs are within the same range at both sites (Table 2;
only high for carbonate-richer samples (R2 = 0.74, n = 35; Fig. 6b). Fig. 7). Notably, even lowest trace element/Al ratios are still (strongly)
Average EFs for Mn at Sites 1259 and 1261 are ~ 0.36 (up to 1.0) and enriched relative to AS (Fig. 5b–f; AS values indicated by black arrows
~ 0.30 (up to 0.49), respectively, indicating Mn depletion throughout on the X-axis). The GH is characterised by consistently lower trace
the investigated sediment intervals (Table 2; Fig. 7). element/Al ratios (Fig. 5b–f).

3.4.2. Cadmium, molybdenum, nickel, vanadium, zinc 3.5. Iron and sulfur
Throughout the study interval at Site 1261 and the upper part at
Site 1259 (499.60–500.10 mcd) there are cyclic patterns of Cd/Al, Mo/ The Fe/Al ratios of the samples investigated are within a relatively
Al, Ni/Al, V/Al and Zn/Al (Fig. 5b–f). The cycles are better resolved at narrow range at Site 1261 (~ 0.44–0.57), while this range is much
Site 1261 compared to Site 1259, partly because the latter had a lower wider at Site 1259 (~ 0.41–1.68), highest Fe/Al ratios being reached
sedimentation rate during biozone CC14 (Bornemann et al., 2008). within the glauconite horizon (Fig. 8a). Small enrichments of Fe/Al at
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 291

Fig. 4. Cross plots of: (a) CaCO3 to Al, (b) Ti to Al, (c) Zr to Al, (d) K to Al, (e) Mg to Al, (f) S/Al to TOC/Al for Sites 1259 (dots) and 1261 (crosses), with respective linear correlation
coefficients (R2).

Site 1261 are located directly below the carbonate-poor intervals (grey highest C/P values coincide with high values of TOC/Al and S/Al, and
bars). However, there are no such peaks in the Fe (CFB) plot (not lowest C/P values with P/Al peaks.
shown here). At Site 1259, there are two Fe/Al peaks in the central part The sequential phosphate extraction yields information about P
of the interval, but these values are more than doubled in the speciation (Fig. 9a, b). Note that P extractions in both cores were
glauconite horizon. The EF of Fe at Site 1261 is ~0.9 on average, with confined to small stratigraphic intervals. The interval at Site 1261 was
maxima of ~1.1 (Table 2). At Site 1259, a slightly higher EF of ~ 1.3 is chosen as all four P peaks, like all the other elemental records, are
observed, but Fe/Al ratios that are consistently N1.0 are mostly repetitive and of similar magnitude (Fig. 8), and most probably
confined to the GH (maximum values = 3.1; AS value = 0.54; Table 2). document similar geochemical processes and conditions. The upper P
The Fe/S ratios (Fig. 8d) can be used as indicator for the amount of peak at Site 1259 was chosen for the same reason (Fig. 8b), while the
sulfide-bound Fe — the lower the ratio, the more Fe is sulfide-bound. glauconite horizon was selected due to its markedly different
The Fe/S ratio at Site 1261 ranges between 0.4 and 0.9, and highest geochemistry.
values are reached within the carbonate-poor intervals, from where The investigated P peak at Site 1261 (8 samples; Fig. 9a) and the
they gradually decrease upcore (Fig. 8d). Thus, highest Fe/S values upper studied P peak at Site 1259 (7 samples, without GH; Fig. 9b)
coincide with lowest trace metal/Al and TOC/Al values. In the upper have a similar P speciation: Fish bones (Step I) and OM-bound P
part of Site 1259 (499.60–500.10 mcd), Fe/S correlates with the same (Step V) make up a minor fraction of the bulk P pool (b10%) at both
parameters as at Site 1261, but Fe/S values are markedly higher (~ 0.6– sites. Detrital apatite (Step IV) is more significant at Site 1259 (~ 0–26%,
1.4), and within the GH, Fe/S reaches maximum values of 2.2 (Fig. 8d). average of ~ 9%) than at Site 1261 (~ 0.2–9%, average of ~ 5%).
Notably, S/Al ratios (Fig. 8e) correlate strongly with TOC/Al at Site 1261 Authigenic apatite (Step III) is clearly the dominant fraction at Site
(R2 = 0.83; Fig. 4f), while the respective correlation is much weaker at 1259 (~ 17–87%, average of ~ 60%), and is also of major importance at
Site 1259 (R2 = 0.27; Fig. 4f). Site 1261 (~ 16–81%, average of ~43%). Iron-bound P (Step II) is the
dominant fraction at Site 1261 (~ 23–66%, average of ~ 49%), while its
3.6. Phosphorus contribution to the P pool is substantially smaller at Site 1259 (~4–
27%, average of ~ 17%). Within the GH at Site 1259 (6 samples; Fig. 9b),
At Site 1261, consistently low P contents of b0.1 wt.% (75 of 115 the fish bone and OM-bound P fractions are likewise rather
samples in total) and P/Al ratios of b0.1 are found (Fig. 8b). However, unimportant (b7%). The sum of iron-bound P and authigenic apatite
pronounced peaks with P contents of up to 2.5 wt.% and high P/Al (~ 6–35% with ~ 21% on average, and ~36–88% ~with 58% on average,
ratios (up to ~ 0.8) cyclically occur within the carbonate-poor layers. At respectively) constitutes the majority of the P pool. In relation to the
Site 1259, we observe P-poor (b0.1 wt.%; 22 out of 101 samples) and P- other P peaks, detrital apatite is more significant within the GH (~1–
rich layers as well, but their pattern is less regular, P peaks are not as 42%, average of ~ 22%).
distinct, and P/Al ratios are generally higher than at Site 1261 (Fig. 8b).
Especially the GH at Site 1259 (499.60–500.10 mcd) is characterized 3.7. SEM and XRD results
by P contents of N2.0 wt.% (up to 5.2 wt.%), and P/Al ratios of N1.0 (up
to 3.3), with higher values at the base of the GH. The TOC to bulk P The majority of particles of the density fraction N3.0 g/cm3 were
ratios (C/P), which is ~ 106:1 in average marine biomass (Redfield, calcium-fluorapatite (CFA), Fe sulfides (mostly pyrite), Zn sulfides and
1958), are variable at both studied sites. C/P ratios range from 3.0 to barite. The analysed samples were selected from core intervals
301 at Site 1261, and from 0.57 to 372 at Site 1259 (Fig. 8c). Generally, representative for the different geochemical and depositional
292 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

Fig. 5. Chemical composition of ODP Sites 1261, 570.2–571.4 mcd, and 1259, 499.6–500.7 mcd. (a) Mn/Al, (b) Cd/Al, (c) Mo/Al, (d) Ni/Al, (e) V/Al, (f) Zn/Al (all as ppm/%) versus core
depth (mcd). Arrows at X-axis indicate respective element/Al values of average shale (Turekian and Wedepohl, 1961).

conditions, and are considered as representative in terms of presence of greigite and mackinawite next to pyrite. Zinc sulfides
composition, size and morphology of the dominant heavy minerals. were observed in the Zn-rich samples (Site 1259, 499.82 and
Pyrite and other Fe sulfides are frequent in all samples, and occur 499.88 mcd) as idiomorphic stochiometric ZnS crystals of 1–5 µm
as euhedral or cubic crystals or as framboids of ~ 2–20 µm in diameter diameter. Crystals were found either dispersed or in framboid-like
(Fig. 10a). The atomic Fe:S ratios of the Fe sulfides are 0.5–1.0 aggregates of ~ 15 µm diameter (Fig. 10b). Barite is present in most of
according to the applied quantification software, indicating the the samples, but rather dispersed, and was observed in the grain size
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 293

Fig. 6. Cross plots of: CaCO3 to Mn/Al for Sites 1261 (a) and 1259 (b), respectively of carbonate-poorer (dots) and -richer (crosses) samples; (c) Mo/Al to TOC/Al, (d) Ni/Al to TOC/Al, (e)
Mo/Al to S/Al, (f) Ni/Al to TOC/Al for Sites 1259 (dots) and 1261 (crosses), with respective linear correlation coefficients (R2).

range of 2–30 µm. Most barites appear as uneven, angular, massive via X-ray diffraction were calcite (confirming the geochemical
grains with signs of fractures (Fig. 10c), most likely fragments of larger analysis) and clinoptilolite ((Na,K,Ca)2–3Al3(Al,Si)2Si13O36·12(H2O)).
grains that were broken into smaller pieces during sediment grinding.
Apatite grains represent the dominant heavy mineral in the P-rich 4. Discussion
samples, while they are absent in the P-poor ones. They cover a wide
range of grain sizes, from few µm to several 100 µm in diameter. 4.1. Terrigenous versus marine sediment input
Irregular aggregates of smaller crystals are most likely of authigenic
origin (Fig. 10d). However, most apatite particles have a massive, Input of Al, K, Mg, Ti and Zr at Sites 1259 and 1261 was rather low,
compact morphology and partly medium to strong rounding, probably with e.g. Al constituting only ~ 1–4 wt.% of bulk sediment. Persistent
due to transportation (Fig. 10d). lamination and strong correlation of Al with K, Mg, Ti and Zr (R2 N 0.9)
In support of the sequential P extraction, large amounts of CFA at Site 1261 support steady sedimentation without bioturbation or
were detected via structural analysis with X-ray diffraction. Other mass wasting, and probably one single detrital source area. Deposi-
common minerals apart from CFA encountered in all analysed samples tional conditions were different at Site 1259 as indicated by the
element/Al plots (Fig. 3) and weaker relationships between the
detrital proxy elements (Fig. 4). They support repeated interruptions
of sedimentation by mass wasting or winnowing of light particles
Table 2
(Nederbragt et al., 2007) and/or change of the detrital source area at
Maximum and mean elemental enrichment factors (EFs) for the investigated intervals
from ODP Sites 1261 and 1259, Demerara Rise, compared to EFs of Cenomanian–
Turonian Boundary Event (CTBE) black shales from Demerara Rise (1) and various CTBE
location worldwide (2; Brumsack, 2006)

Mean EF of Max. EF of Mean EF of Max. EF of EF of (1) EF of (2)


1261 vs. AS 1261 vs. AS 1259 vs. AS 1259 vs. AS vs. AS vs. AS
Ba/Al 4.78 5.85 3.04 6.39 2.35 3.10
Cd/Al 212 763 298 1004 170 184
Fe/Al 0.88 1.03 1.28 3.14 1.16 1.44
K/Al 0.96 1.03 1.06 1.85 0.86 0.93
Mg/Al 1.12 1.37 1.35 2.49 1.41 1.29
Mn/Al 0.30 0.49 0.36 1.10 0.44 1.29
Mo/Al 130 207 74.2 164 127 208
Ni/Al 10.4 16.1 8.74 13.7 10.6 11.5
P/Al 13.5 71.8 67.2 418 8.80 5.91
S/Al 30.9 38.8 27.5 46.5 – –
Ti/Al 0.83 0.85 0.84 0.85 1.01 0.96
V/Al 30.7 46.8 32.7 46.8 33.2 18.4
Fig. 7. Bar plots displaying mean elemental enrichment factors (EFs; note logarithmic
Zn/Al 30.9 86.7 35.1 86.7 22.8 42.5
scale) of Sites 1259 and 1261, as well as of OAE2 black shale from Demerara Rise
Zr/Al 0.67 0.80 0.70 0.80 0.83 0.99
(Brumsack, 2006) and OAE2 mean (from various OAE2 locations by Brumsack, 2006),
All EFs shown are calculated from element/Al ratios relative to average shale (AS; relative to average shale (Turekian and Wedepohl, 1961). Dashed lines indicate EFs of 5,
Turekian and Wedepohl, 1961). 30 and 100.
294 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

Fig. 8. Chemical composition of ODP Sites 1261, 570.2–571.4 mcd, and 1259, 499.6–500.7 mcd. (a) Fe/Al, (b) P/Al, (c) C/P, (d) Fe/S, (e) S/Al (all as %/%) versus core depth (mcd). Boxes in
(b) mark the intervals of sequential P extraction. The dashed lines (c) indicate the C/P ratio of average marine organic plankton (after Redfield, 1958).

Site 1259. A possible explanation for the small amount of continental constitutes ~30–70 wt.% of bulk sediment at both sites, supporting a
material reaching the relatively proximal Sites 1259 and 1261 (Fig. 1c) strong influence of carbonate productivity on sedimentation. X-ray
could be the trapping of most detrital matter on the flooded shelf area diffraction data from both sites reveal the presence of large amounts
due to high sea level (Mosher et al., 2007; Nederbragt et al., 2007) in of clinoptilolite. Clinoptilolite is one of the most common zeolite
the upper Cretaceous. minerals in marine pelagic sediments, and has previously been
The observed cyclicity of the Al record, particularly at Site 1261, is described in Cretaceous black shales from the North Atlantic (Thurow,
attributed to variable dilution of the detrital signal by CaCO3 which 1988). It is formed authigenically by dissolution of amorphous SiO2
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 295

Fig. 9. Cummulative extracted P fractions at (a) Site 1261 (570.56–570.64 mcd) and (b) Site 1259 (500.1–500.53 mcd) in g PO4/kg sediment versus core depth (mcd). Explanation of
extraction steps in Table 1. Note that compared to other geochemical data presented, P extraction was only applied to comparably small core intervals.

(volcanic glass, opal tests as Si source) during burial diagenesis (e.g. The interpretation of redox-sensitive (trace) metals investigated at
Kastner, 1980). In Demerara Rise black shales, preservation of siliceous Demerara Rise will be preceded by a short introduction into its
radiolarian tests is generally poor (Erbacher et al., 2004; Musavu- behaviour under variable redox conditions.
Moussavou and Danelian, 2006; Nederbragt et al., 2007), most
probably related to extensive silica diagenesis in these deeply buried 4.2.1. Manganese
sediments (e.g. Kastner, 1980; Thurow, 1988). Karpoff et al. (2007) In general, Mn is reductively leached from marine sediments under
regarded clinoptilolite as a diagnostic proxy for paleoproductivity in a sub- to anoxic conditions (e.g. Froelich et al., 1979; Klinkhammer and
Miocene marine setting. We therefore assume that clinoptilolite in the Bender, 1980; Lewis and Landing, 1991; Burdige, 1993; Morford and
studied samples is derived from dissolution-reprecipitation reactions Emerson, 1999). In a stratified anoxic water body, Mn2+/Mn (oxyhydr)
of biogenic silica, and thus indicative of opal productivity at the oxide cycling can occur shortly above the oxic/anoxic transition (e.g.
Coniacian–Santonian Demerara Rise. In further support for elevated Black Sea, Cariaco Basin; Tebo, 1991; Yakushev et al., 2007; Percy et al.,
marine primary paleoproductivity, organic geochemical studies in press), and may thus scavenge and shuttle other trace metals to the
within Cretaceous black shales at Demerara Rise (e.g. Meyers et al., deeper, anoxic water column (e.g. Morford et al., 2005; Brumsack,
2006; Forster et al., 2007b; Beckmann et al., in press) proved an overall 2006; Tribovillard et al., 2006). In an open margin oxygen minimum
high contribution of fresh marine OM to the bulk sediment. zone (OMZ), Mn2+ leached from the underlying sediment may partly
Taking all these results into account we infer a depositional regime diffuse out of the OMZ (e.g. Klinkhammer and Bender, 1980; Lewis and
at both ODP Sites that was dominated by strong biogenic marine input. Luther, 2000; Brumsack, 2006), resulting in sedimentary Mn deple-
This conclusion is consistent with interpretations for OAE3 black shales tion. The significant Mn depletion within the studied OAE3 intervals
from the Deep Ivory Basin (ODP Site 959) and OAE2 black shale from (except for the lowest part of the 1259 interval) indicates sedimentary
the Tarfaya shelf basin off SW Morocco (e.g. Wagner, 2002; Wagner Mn leaching under constantly sub-/anoxic conditions (Fig. 5a).
et al., 2004; Beckmann et al., 2005a,b). Estimates of paleoproductivity However, the low amounts of sedimentary Mn appear to be bound
based on the barite record are hampered by diagenesis (Brumsack, to carbonate as indicated by a strong correlation between carbonate
2006). Pore water profiles of SO42− and Ba2+ (Erbacher et al., 2004) and Mn/Al at least for carbonate-rich samples (Fig. 6a, b). By XRD
indicate sulfate depletion throughout the black shales at Demerara analyses, no Mn carbonate was detected, probably due to its low
Rise, resulting in barite dissolution in this interval for the last ~ 86 Ma contents (b1 wt.%). The Mn to carbonate coupling is most obvious in
(Arndt et al., 2006). Despite overall Ba enrichment relative to AS the bottom part of the 1259 section (below 500.48 mcd; Fig. 4a), where
(Fig. 7), SEM analyses of samples from ODP Sites 1259 and 1261 did not both Mn/Al and carbonate reach maximum values, implying sub-
identify any biogenic barite crystals, which are generally smaller than stantial Mn carbonate (rhodochrosite) formation. Minor differences in
diagenetic ones (0.5–5 µm versus 20–700 µm), and form elliptical or the degree of depletion at both sites may be due to higher
euhedral crystals or aggregates (Torres et al., 1996; Paytan et al., 2004; sedimentation rates at Site 1261 (resulting in higher time resolution
Paytan and Griffith, 2007). Instead, we detected numerous (fragments of the sample set). Alternatively, physical sedimentation processes at
of) most probably diagenetic barite particles (Fig. 10c). Site 1259 might have disturbed the carbonate–Mn correlation. The
reason for different slopes of Mn and carbonate correlation (Fig. 6a, b)
4.2. Redox-sensitive elements could be due to variations in the contribution of rhodochrosite to bulk
carbonate. More precisely, the best fit correlation line for carbonate-
Combining records of different redox-sensitive (trace) elements is richer samples at Site 1261 has a positive intercept with the carbonate
a well established approach for reconstructing past redox conditions. axis (Fig. 6a), indicating that other carbonate phases are present in
296 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 297

these samples. The respective line for carbonate-poorer samples sink (Huerta-Diaz and Morse, 1992; Dellwig et al., 2002; Algeo and
intercepts with the Mn/Al axis, indicating that Mn is not only bound in Maynard, 2004). In addition, during OM sulfurization or formation of
carbonates, but probably also to silicates or even oxides. The formation geoporphyrins a secondary coupling of Ni and Mo to OM can be
of Mn carbonates is most probably a diagenetic process close to the created (e.g. Lewan and Maynard, 1982; Breit and Wanty, 1991; Algeo
sediment surface, where OM degradation via Mn (oxyhydr)oxide and Maynard, 2004; Tribovillard et al., 2004). The distinct correlations
reduction increases pore water alkalinity and Mn2+ (e.g. Baltic Sea; of Mo/Al and Ni/Al not only with TOC/Al, but also with S/Al at Sites
Huckriede and Meischner, 1996; Neumann et al., 2002). However, 1259 and 1261 (Fig. 6c–f) imply that these trace metals were mainly
carbonate–Mn correlation might also be obtained syngenetically by incorporated into (sulfurized) OM during early diagenesis. März et al.
Mn–carbonate overgrowths on carbonate tests within Mn2+-enriched (in press) showed that at Site 1261, only ~ 1/3 of bulk S is bound to
anoxic water bodies (Brumsack, 2006), which could have been the case metal sulfides, indicating a significant contribution of sulfurized OM
during biozone CC14 at Sites 1259 and 1261. Still, relatively elevated to the total S pool. A close coupling of Mo to sulfurized OM was found
Mn/Al and carbonate contents at the bottom of 1259 might indicate in several black shale deposits (Tribovillard et al., 2004; Algeo and
intermittent and minor influence of oxygen and diagenetic rhodo- Lyons, 2006). In contrast, we assume that Ni was rather incorporated
chrosite precipitation. Precipitation of Mn sulfides is generally into organic tissue as a micronutrient already during primary
neglected as Mn sink in black shales, as is Mn incorporation into production and is not as strongly related to OM sulfurization as Mo.
pyrite or organic phases (Huerta-Diaz and Morse, 1992; Algeo and This is indicated by slightly stronger correlation of Ni/Al with TOC/Al
Maynard, 2004). than with S/Al at both sites, while Mo/Al is correlated with TOC/Al and
S/Al to the same degree. The difference towards Cd and Zn is most
4.2.2. Cadmium and zinc probably an effect of higher HS− concentrations being required for Mo
These elements are present under oxic to suboxic conditions as and Ni coupling to OM. These can only be reached when the initial
soluble cations (CdCl+; Zn2+, ZnCl+; Tribovillard et al., 2004) or are “buffering” of low HS− in bottom waters by CdS and ZnS formation is
incorporated into marine plankton biomass, which is their major declining.
carrier to the sea floor (e.g. Bruland et al., 1991). Already at very low
HS− concentrations, Cd and Zn may form sulfides (CdS, ZnS), due to 4.2.4. Vanadium
lower solubility products and faster water exchange kinetics than e.g. Under oxic to suboxic conditions, V is present as an oxy-anion
Fe sulfides (Jacobs and Emerson, 1982; Rosenthal et al., 1995; Morse (HVO42−, H2VO4−). Manganese (oxyhydr)oxides can play a significant
and Luther, 1999; Scholz and Neumann, 2007). Indeed, in the studied role for its transport to the sea floor (e.g. Wehrli and Stumm, 1989;
sediments, sharp Zn and Cd peaks (Fig. 5b, f) document rapid and Hastings et al., 1996). Via a two-step reaction pathway, V forms
intensive precipitation of CdS and ZnS in weakly sulfidic bottom hydroxyl species (VO(OH)3−) and insoluble hydroxides (VO(OH)2)
waters (Morse and Luther, 1999; März et al., in press). This is under anoxic, and insoluble hydroxides (V(OH)3) and oxides (V2O3)
supported by SEM-EDX detection of pure idiomorphic ZnS crystals under sulfidic conditions (e.g. Breit and Wanty, 1991; Tribovillard
(Fig. 10b). Above the Cd/Al and Zn/Al peaks, values decline rapidly, et al., 2004). Incorporation of V into pyrite under sulfidic conditions is
most probably due to a drawdown of the respective trace metal pools regarded as negligible (e.g. Dellwig et al., 2002; Scholz and Neumann,
in the ambient water masses (Algeo, 2004). It is hard to estimate the 2007), but V is incorporated into geoporphyrins (Lewan and Maynard,
spacial extension of this drawdown, but as it is documented by our 1982; Breit and Wanty, 1991), and thus coupled to OM similar to Ni. At
data at both ODP Sites 1259 and 1261, this phenomenon at least Sites 1259 and 1261, the V/Al records (Fig. 5e) show patterns similar to
affected larger parts of the Demerara Rise continental slope. In Cd/Al and Zn/Al, but with a more gradual decline of V/Al above the
samples with high Zn contents, also Fe sulfides were observed which initial peak. Increasing incorporation of V into geoporphyrins could be
do not contain Cd and Zn, supporting negligable incorporation of responsible for keeping V/Al values relatively high, while Cd/Al and
these trace elements into pyrite (Huerta-Diaz and Morse, 1992; Scholz Zn/Al rapidly declined (Breit and Wanty, 1991; Algeo and Maynard,
and Neumann, 2007). Weak correlations of Cd/Al and Zn/Al with both 2004). Alternatively, bottom water V was maybe not exhausted as fast
TOC/Al and S/Al at both sites (correlation coefficients of 0.01–0.36) as the Cd and Zn pools. Initial V (hydr)oxide formation was probably
indicate that no significant fraction of sedimentary Cd and Zn is bound limiting sedimentary Ni enrichment, as the reduction from V(IV) to V
to (sulfurized) OM or pyrite. (III) at low HS− concentrations is known to inhibit Ni incorporation
into geoporphyrins (Lewan, 1984; Breit and Wanty, 1991).
4.2.3. Molybdenum and nickel
These elements form soluble cations (Ni2+, NiCl+) or oxy-anions 4.2.5. Iron
(MoO42−) under oxic to suboxic conditions. Nickel is a micronutrient Under oxygen- and nitrate-free conditions, Fe (oxyhydr)oxides are
taken up by primary producers and transported to the sea floor mainly progressively reduced (Froelich et al., 1979; review by Burdige, 1993),
by OM (e.g. Bruland et al., 1991). Although conservative in ocean water either chemically (e.g. Stone and Morgan, 1987; Stumm and
under most conditions, Mo is important for biological N fixation (e.g. Sulzberger, 1992) or biologically (dissimilatory iron reduction; e.g.
Tuit et al., 2004) and may be coupled to Mn (oxyhydr)oxides, which Lovley, 1987; Nealson and Myers, 1992), and Fe2+ is liberated to the
then act as Mo shuttles to the sea floor under non-sulfidic conditions surrounding (pore) water. If free HS− is present, Fe (oxyhydr)oxides are
(e.g. Shimmield and Price, 1986; Adelson et al., 2001; Reitz et al., reduced (e.g. Morse et al., 1987; Yao and Millero, 1996) and
2007). In response to algal blooms, a transport of Mo to the sea floor transformed to Fe sulfides, mostly pyrite (FeS2; e.g. Berner, 1984;
via organic macromolecules forming in the water column has been Schoonen, 2004). However, if the pool of most sulfide-reactive Fe
postulated (e.g. Lunau et al., 2006; Dellwig et al., 2007). In sulfidic minerals (e.g. ferrihydrite, goethite, hematite; Poulton et al., 2004) is
environments, Mo oxy-anions are transformed to particle-reactive exhausted, HS− may start to react with fresh OM in the sediment (OM
thiomolybdates above a certain HS− threshold (“thiomolybdate sulfurization, natural vulcanisation; e.g. Sinninghe Damsté and de
switch”; Helz et al., 1996; Zheng et al., 2000; Adelson et al., 2001), Leeuw, 1990; Sinninghe Damsté et al., 1998). Excess HS− may also
and scavenged from the water column by e.g. sinking OM. In sulfidic diffuse out of the sediment, creating sulfidic conditions at the sea floor
sediments, incorporation into pyrite can be an important Mo and Ni and in the (lower) water column. Sequential Fe extraction results

Fig. 10. SEM micrographs in back scatter electron mode and respective elemental spectra from representative particles. (a) Fe sulfide framboids (1–5); (b) aggregate/framboid of
idiomorphic Zn-sulfide crystals (1–4); (c) angular barite grain, probably fragment of larger diagenetic barite aggregates; (d) aggregate of fine apatite crystals (1–4), Fe sulfide with clay
coating (5). EDS spectra numbers correspond to the spot analyses of the respective numbered particles.
298 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

obtained in an earlier study (März et al., in press) indicated repeated sites (including Site 1261, where black shales started to develop before
shifts from sulfidic to anoxic, non-sulfidic bottom water conditions at the OAE2), implying that during the Cretaceous Site 1259 was situated
Site 1261. Similar information can be obtained from the sedimentary in shallower water depths than Site 1261. Benthic foraminiferal
Fe/S ratios, available for both Sites (Fig. 8d). Periodic increases in Fe/S assemblages also support that re-oxygenation of the Demerara Rise
ratios at Sites 1259 and 1261 indicate higher amounts of non-sulfide- sea floor occurred first at Site 1259 followed by the other ODP Leg 207
bound sedimentary Fe. Notably, Fe/S peaks do not coincide with sites (Friedrich and Erbacher, 2006).
elevated Fe/Al values at Site 1261 (Fig. 8a), implying that these Fe/Al
peaks do not represent higher amounts of Fe oxides, but probably 4.3.1. Sediment source
formation of syngenetic pyrite in a sulfidic water column. Co- From the high-resolution profiles presented in this study, it
occurrence of Fe/S peaks with lowest–but compared to AS still appears that the cyclicity of the carbonate, Al and TOC/Al records,
enriched–trace element/Al, lowest TOC/Al, and highest P/Al values and thus the overall sedimentary input is much more regular at Site
points to less reducing, but still anoxic conditions. High Fe/S and Fe/Al 1261 compared to Site 1259. Based on Al and carbonate data, we infer
ratios within the GH are probably a combined effect of elevated Fe- that variable dilution by marine carbonate was the main process
oxide (6–35% of total P bound to Fe-oxides) and glauconite (containing leading to the observed fluctuations in carbonate, Al and probably
~ 20% Fe) contents. As stated above, März et al. (in press) found that TOC/Al. Well-preserved mm-scale lamination at Site 1261 excludes
only about 15–35% of bulk S at Site 1261 are bound to pyrite (see also physical and biological disturbance of the sediment in the investigated
Böttcher et al., 2006), indicating that sulfurized OM may be a major S interval. The strong correlation between all terrigenous “marker
carrier in these black shales. elements” such as Al, K, Mg, Ti and Zr (Fig. 4a–e) implies a stable
detrital source area at Site 1261. “Terrestrial” elements (Fe, K, Mg, Ti,
4.2.6. Phosphorus Zr) normalized to Al support that the lithology of the detrital source
In the ocean, P is mainly present as HPO42−, which is not redox- material is close to AS composition (Fig. 7), consistent with the
sensitive itself, but strongly coupled to the cycles of Fe and OM. Fresh findings of Hetzel et al. (2006). At Site 1259, however, the terrestrial-
iron (oxyhydr)oxides cannot only adsorb large amounts of phosphate sourced elements are overall slightly more enriched than at Site 1261
ions, but phosphate may also be co-precipitated during Fe (oxyhydr) (Table 2; Fig. 7), implying a stronger detrital influence, and thus a
oxide formation. Thus, Fe (oxyhydr)oxides can play a major role in the more proximal location of Site 1259 during the Coniacian–Santonian
transport of P to the sea floor under oxic conditions (e.g. Feely et al., consistent with Erbacher et al. (2005) and Friedrich and Erbacher
1990; Slomp et al., 1996; Poulton and Canfield, 2006). Another P (2006). Alternatively, the input of marine biogenic carbonate could
carrier to the sea floor is marine OM, as marine phytoplankton have been lower at Site 1259. Weaker carbonate-Al anti-correlation,
contains a relatively well-defined amount of P (mean C/P ratio of weaker correlations of the detrital elements and cyclic Zr/Al variations
106:1; Redfield, 1958). Under reducing conditions, a gradual decou- support that the interplay of marine versus terrigenous dilution was
pling of P from both the Fe and the OM cycles takes place: Dissolution disrupted by mass wasting, winnowing of fine particles, and/or
of Fe (oxyhydr)oxides liberates phosphate to the surrounding water. varying sources of lithogenic input. The periodic, potentially erosional
Furthermore, under anoxic conditions there is a preferential regen- activity of winnowing bottom currents would also offer an explana-
eration of organic P from OM in relation to organic carbon (e.g. Ingall tion for lower sedimentation rates at Site 1259 than at Site 1261,
et al., 1993; Ingall and Jahnke, 1997), resulting in low P burial where continuous hemipelagic sedimentation is documented. The
capacities of anoxic sediments and C/P ratios of N106 (e.g. Anderson occurrence of glauconitic horizons, as observed at Site 1259, is
et al., 2001; review by Algeo and Ingall, 2007). Our bulk sediment P interpreted by Friedrich and Erbacher (2006) and Nederbragt et al.
data indicate P depletion over large parts of the investigated intervals, (2007) as an indication for extremely low sedimentation rates,
resulting in high C/P and low P/Al ratios (Fig. 8b, c). This implies a very possibly periodic erosion due to increased bottom currents at Site
effective regeneration of both Fe (oxyhydr)oxide- and OM-bound P 1259. Indeed, coarser grain size, enrichment of heavy mineral-bound
from the sediment under anoxic, probably even sulfidic bottom water Zr (Fig. 3d), association of glauconite and detrital apatite (identified by
conditions. However, P/Al enrichments at both sites, with very low C/P sequential P extraction and SEM analysis; Figs. 9a and 10d), and low
ratios and Fe-bound P as one of the major fractions (Figs. 8b, c and 9), carbonate contents within the GH support this conclusion. Still,
document conditions favorable for P burial, and even (partial) enrichments of K/Al, Mg/Al, and depletion of Ti/Al are no typical
preservation of Fe-oxides (and the Fe–P coupling), in the sediment. features of a winnowing horizon, but rather suggest an import of
Partial transfer of Fe-bound P, and nearly complete transfer of OM- detrital material from a different source area with a K- and Mg-richer,
bound P, to authigenic apatite is consistent with the “sink switching” and Ti-poorer lithology. Unfortunately we cannot be more precise in
hypothesis, i.e. P transfer from rather labile binding forms into that respect due to limited knowledge about the detrital source area or
authigenic apatite as the ultimate oceanic P sink (e.g. Delaney, 1998; related drainage systems to Demerara Rise.
Filippelli, 2001). The high P content of the GH is probably a mixed
signal of originally larger amounts of Fe oxide-bound P transformed to 4.3.2. Redox cycles
the now dominant authigenic CFA, and input of older, well-crystal- Rapid and cyclic bottom water redox fluctuations at Site 1261 were
lized authigenic and/or detrital apatite (Fig. 9b). The fact that apatite suggested to document variations between sulfidic and anoxic, non-
grains identified under the SEM in GH samples are often rounded sulfidic conditions in an earlier work of März et al. (in press).
(Fig. 10d) points to re-deposition, e.g. by bottom currents. Comparing the records of Cd/Al, Mo/Al, Ni/Al, V/Al and Zn/Al at Site
1261 and the part of 1259 above the GH, the variations from maxima
4.3. Geochemistry at sites 1259 and 1261 — similarities and differences to minima (Fig. 5b–f) and the respective trace element EFs (Fig. 7) are
similar at both sites, despite lower time resolution at Site 1259. This
Comparing quasi age-equivalent records from the two drill sites on indicates that at least the upper part of Site 1259 (499.60–500.10 mcd)
Demerara Rise in terms of a paleo-water depth transect bears a experienced similar redox variations as Site 1261. Detection of
fundamental complication. Cretaceous water depths were most variable, but overall anoxic conditions not only at the supposedly
probably different than today due to subsidence and tilting of the shallower Site 1259, but also at the deeper Site 1261 suggests that
Demerara Rise during progressive rifting in the South Atlantic since indeed large parts of the continental margin at Demerara Rise were
the earliest Cretaceous (Erbacher et al., 2004). In this context, Erbacher affected by bottom water anoxia during OAE3. However, more detailed
et al. (2005) stated that transgressive Cretaceous black shale comparison of redox-sensitive proxy records reveals slight differences
deposition at Site 1259 started later than at all other ODP Leg 207 between both sites. Compared to the four nearly identical redox cycles
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 299

at Site 1261 (regular cyclicity, very similar maximum and minimum During OAE2, Hetzel et al. (2009) find a significant depletion of
element/Al values in each cycle, repetitive “sawtooth” patterns), Site Zn (with mean Zn/Al values of ~ 50 ⁎ 10− 4 and ~155 ⁎ 10− 4 at Sites 1258
1259 displays more irregular variations, with some element/Al and 1260, respectively) compared to the under- and overlying
minima (500.15–500.22 mcd; Fig. 5b–f) and maxima (499.92– sediments. This is in contrast with the variable, but generally elevated
499.98 mcd; Fig. 5b–f) that are less pronounced than expected. Zn/Al values during the studied OAE3 interval (mean Zn/Al values of
Thus, we conclude that although redox cyclicity at Site 1259 above the ~380 ⁎ 10− 4 and ~ 335 ⁎ 10− 4 at Sites 1259 and 1261, respectively;
GH was similar to Site 1261, the intensity of anoxia/euxinia and the Fig. 5f). However, we observe repeated intervals with lower Zn/Al
regularity of redox changes were less pronounced. A major difference values, comparable to the mean OAE2 values of Hetzel et al. (2009).
between both sites is the occurrence of the glauconite horizon at Site As most probable explanation for these unexpected low Zn/Al in black
1259 (dotted bar, Figs. 2, 3, 5, 8). The relatively low Al ratios of Cd, Ni, shales, Hetzel et al. (2009) state a global drawdown of the Zn sea water
Mo, V and Zn throughout the GH (Fig. 5b–f), in combination with pool, resulting from extreme and wide-spread ocean anoxia to euxinia
lower TOC/Al (Fig. 2b) and high Fe/S (Fig. 8d) ratios point to less and subsequent syn- or diagenetic ZnS formation. This hypothesis is
reducing conditions during its deposition than during the overlying consistent with the model of März et al. (in press), where repeated
section of Site 1259. In support of this, the P/Al and C/P records within establishment of weakly sulfidic bottom water conditions leads to
the GH show highest (0.6–3.3) respectively lowest (0.3–2.9) values of rapid precipitation of CdS and ZnS (Fig. 5b, f), followed by (at least
all samples (Fig. 8b, c), proving enhanced P retention in the sediment local) drawdown of the Cd and Zn sea water inventories. We therefore
under at least non-sulfidic conditions. In addition, ~ 20% of all P suggest that similar processes of trace metal drawdown were active
measured in the GH is Fe oxide-associated (Fig. 9a). We assume that during both OAE2 and 3. Still, while having a global impact on oceanic
depositional conditions during formation of the GH were probably trace metal budgets during OAE2, their effects were less persistant
influenced by enhanced bottom current activity, introducing less during OAE3 and probably restricted to extreme environments like
reducing bottom waters and winnowing fine particles. Demerara Rise. This interpretation is consistent with the C and N
The sediment section below the GH at Site 1259 (below isotope data of Junium and Arthur (2007). They assume similar
500.48 mcd) is the only interval where Mn/Al reaches values geochemical conditions in the water column during both OAEs, yet
comparable to AS (Fig. 5a). This Mn/Al maximum is paralleled by describe OAE2 as a global anoxic event, while OAE3 affected the global
highest overall carbonate contents of up to ~100 wt.% (Fig. 2c). As carbon and nitrogen cycles only locally.
discussed earlier, Mn carbonate formation requires elevated Mn2+
levels in the pore water, which are induced by reductive dissolution of 5. Conclusion
Mn (oxyhydr)oxides. Deposition of Mn (oxyhydr)oxides requires
periodically oxic bottom water conditions. Indeed, less reducing During Coniacian–Santonian time ODP Sites 1261 and 1259 were
bottom water conditions are implied by relatively low Cd/Al and Zn/ dominated by marine (calcareous and opal) sedimentation and burial
Al ratios (Fig. 5b, f). Also low C/P ratios indicate P burial in the sediment of large amounts of OM, which at least partly related to elevated OM
under oxic conditions (Fig. 8c). However, V/Al and especially Mo/Al and preservation by diagenetic OM sulfurization. The presented new
Ni/Al exhibit very high values (Fig. 5c–e), opposing oxic conditions. geochemical data support the conclusion that both sites were affected
Notably, the trace elements highly enriched below the GH are those by similar bottom water redox conditions throughout the investigated
associated to OM (which is in support of the high TOC/Al values in the intervals, which were generally anoxic, and even sulfidic for prolonged
respective interval), while sulfide-bound Cd and Zn are relatively low. periods. However, bottom water and sediment redox were subject to
Despite these two contradictory interpretations for the interval below relatively rapid and cyclic variations, which occurred at both sites. This
the GH, its redox development did obviously not follow the regular observation suggests that prevailing redox conditions affected large
systematics as the redox cycles above the GH and the ones at Site 1261. parts of the continental margin at Demerara Rise during OAE3. The
sedimentary record documents that redox fluctuations were regular
4.3.3. Comparison with OAE2 at Demerara Rise in frequency and degree of changes at the deeper Site 1261. At the
Comparison of OAE3 deposits from Sites 1259 and 1261 with shallower position of Site 1259 sedimentation instead was stronger
values reported for the OAE2 (Cenomanian–Turonian) at Demerara affected by variations in detrital supply and/or winnowing of bottom
Rise (Brumsack, 2006; Table 2; Fig. 7) shows very similar EFs for most currents, and probably periodic erosion, together supporting a more
trace elements. A direct comparison of OAE3 high-frequency redox proximal location. In contrast, steady hemipelagic sedimentation
cycles is hampered by the lack of high-resolution records for OAE2. prevailed at Site 1261, consistent with a greater water depth.
However, a geochemical study of sediments deposited before, during Especially during the formation of the glauconitic horizon at Site
and after OAE2 at ODP Sites 1258 and 1260 (Hetzel et al., 2009) 1259, the depositional regime was physically disturbed, and bottom
allows some comparison of geochemical processes occurring at waters were most probably better oxygenated. Speciation of P
Demerara Rise during both OAEs. indicates that at both sites Fe oxide-bound P and authigenic apatite
Similar to OAE3, Hetzel et al. (2009) state that organic matter is are the dominant sedimentary P species, while detrital apatite seems
an important carrier of S in OAE2 deposits at Demerara Rise, which is to be of higher importance within the GH. Only minor P contributions
due to low input of sulfide-reactive iron to the sedimentary system. are related to fish bones and organic P. Geochemically, the studied
Also Böttcher et al. (2006) found similar Fe–S systematics in OAE2 OAE3 deposits are well-comparable to the OAE2 black shales from
black shales at Demerara Rise, indicating an Fe-limited system, a high Demerara Rise, indicating that both anoxic events created very similar
degree of pyritization and intensive OM sulfurization. Obviously, paleoenvironmental conditions at this location, although the impact
detrital iron delivery to Demerara Rise was low during both OAEs, of OAE3 on marine biogeochemical cycles was probably rather local.
while HS− production was comparably high.
Manganese is generally depleted in sediments from both OAEs as Acknowledgements
well (Figs. 5a, 7), which is attributed to its reductive leaching from the
sediment under anoxic conditions and removal in an open-margin We thank W. Hale for sampling assistance, and B. Kockisch, S. Pape,
oxygen minimum zone. Intervals with relatively elevated Mn/Al S. Siemer and K. Enneking for analytical and logistic support. Help of K.
values occur only in association with very carbonate-rich layers during Küster, J. Hoffmann and N. Allroggen with the phosphate extraction is
both OAEs, which is not necessarily associated with oxidation events, highly appreciated. Electron microscopy was performed at the EMU
but might also be related to diagenetic formation of carbonate layers (Utrecht University), and we thank M. Drury for his support. CM
or sediment layers with reduced porosity (Hetzel et al., 2009). thanks G. de Lange and the Geochemistry group at Utrecht University
300 C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301

for great hospitality during his research stay. TW acknowledges the Flögel, S., Wagner, T., 2006. Insolation-control on the Late Cretaceous hydrological cycle
and tropical African climate — global climate modelling linked to marine climate
Royal Society-Wolfson Research Merit Award. This study was funded records. Palaeogeogr. Palaeoclimatol. Palaeoecol. 235, 288–304.
by the DFG via the International Graduate College EUROPROX. The Flögel, S., Beckmann, B., Hofmann, P., Bornemann, A., Norris, R.D., Wagner, T., in press.
manuscript benefited significantly from constructive comments of A. Evolution of tropical watersheds and continental hydrology during the Late
Cretaceous greenhouse; marine carbon burial and possible implications for the
Negri and two anonymous reviewers. future. Earth Planet. Sci. Lett.
Forster, A., Schouten, S., Moriya, K., Wilson, P.A., Sinninghe Damsté, J.S., 2007a. Tropical
warming and intermittent cooling during the Cenomanian/Turonian oceanic anoxic
References event 2: sea surface temperature records from the equatorial Atlantic. Paleoceano-
graphy 22, PA1219. doi:10.1029/2006PA001349.
Adelson, J.M., Helz, G.R., Miller, C.V., 2001. Reconstructing the rise of recent coastal Forster, A., Schouten, S., Baas, M., Sinnighe Damsté, J.S., 2007b. Mid-Cretaceous (Albian–
anoxia: molybdenum in Chesapeake Bay sediments. Geochim. Cosmochim. Acta 65, Santonian) sea surface temperature records of the tropical Atlantic Ocean. Geology
237–252. 35, 919–922.
Algeo, T.J., 2004. Can marine anoxic events draw down the trace-element inventory of François, R., 1988. A study on the regulation of the concentrations of some trace metals
seawater? Geology 32, 1057–1060. (Rb, Sr, Zn, Pb, Cu, V, Cr, Ni, Mn and Mo) in Saanich Inlet sediments, British
Algeo, T.J., Maynard, J.B., 2004. Trace element behaviour and redox facies in core shales Columbia, Canada. Mar. Geol. 83, 285–308.
of the Upper Pennsylvanian Kansas-type cyclothems. Chem. Geol. 206, 289–318. Franke, C., Frederichs, T., Dekkers, M.J., 2007. Efficiency of heavy liquid separation to
Algeo, T.J., Lyons, T.W., 2006. Mo-total organic carbon covariation in modern anoxic concentrate magnetic particles. Geophys. J. Int. doi:10.1111/j.1365- 246X.2007.03489.
marine environments: implications for analysis of paleoredox and paleohydro- Friedrich, O., Erbacher, J., 2006. Benthic foraminiferal assemblages from Demerara Rise
graphic conditions. Paleoceanography 21, PA 1016. (ODP Leg 207, western tropical Atlantic): possible evidence for a progressive
Algeo, T.J., Ingall, E., 2007. Sedimentary Corg:P ratios, paleocean ventilation, and opening of the Equatorial Atlantic Gateway. Cretac. Res. 27, 377–397.
Phanerozoic pO2. Palaeogeogr. Palaeoclimatol. Palaeoecol. 256, 130–155. Friedrich, O., Erbacher, J., Mutterlose, J., 2006. Paleoenvironmental changes across the
Anderson, L.D., Delaney, M.L., Faul, K.L., 2001. Carbon to phosphorus ratios in sediments: Cenomanian/Turonian boundary event (oceanic anoxic event 2) as indicated by
implications for nutrient cycling. Glob. Biogeochem. Cycles 15, 65–79. benthic foraminifera from the Demerara Rise (ODP Leg 207). Rev. Micropaleontol.
Arndt, S., Brumsack, H.-J., Wirtz, K., 2006. Cretaceous black shales as active bioreactors: 49, 121–139.
a biogeochemical model for the deep biosphere encountered during ODP Leg 207 Froelich, P.N., Klinkhammer, G.P., Bender, M.L., Luetge, N.A., Heath, G.R., Cullen, D.,
(Demerara Rise). Geochim. Cosmochim. Acta 70, 480-425. Dauphin, P., 1979. Early oxidation of organic matter in pelagic sediments of the
Beckmann, B., Flögel, S., Hofmann, P., Schulz, M., Wagner, T., 2005a. Orbital forcing of eastern equatorial Atlantic: suboxic diagenesis. Geochim. Cosmochim. Acta 43,
Cretaceous river discharge in tropical Africa and ocean response. Nature 437, 1075–1090.
241–244. Goldstein, J.I., Newbury, D.E., Echlin, P., Joy, D.C., Romig Jr., A.D., Lyman, C.E., Fiori, C.,
Beckmann, B., Wagner, T., Hofmann, P., 2005b. Linking Coniacian–Santonian (OAE3) Lifshin, E., 1992. Scanning electron microscopy and X-Ray microanalysis, 2nd
black shale formation to African climate variability: a reference section from the edition. Plenum Press, New York. 820 pp.
eastern tropical Atlantic at orbital time scales (ODP Site 959, off Ivory Coast/ Ghana). Hardas, P., Mutterlose, J., 2006. Calcareous nannofossil biostratigraphy of the
In: Harris, N.B. (Ed.), The Deposition of Organic Carbon-Rich Sediments: Models, Cenomanian/Turonian boundary interval of Leg 207 at the Demerara Rise. Rev.
Mechanisms, and Consequences. SEPM Spec. Publ., vol. 82, pp. 125–143. Micropaleontol. 49, 165–179.
Beckmann, B., Hofmann, P., März, C., Schouten, S., Sinninghe Damsté, J.S., Wagner, T., in Hastings, D.W., Emerson, S.R., Mix, A.C., 1996. Vanadium in foraminiferal calcite as a
press. Coniacian-Santonian deep ocean anoxia/euxinia inferred from molecular and tracer for changes in the areal extent of reducing sediments. Paleoceanography 11,
inorganic markers: Results from the Demerara Rise (ODP Leg 207). Org. Geochem. 665–678.
Berner, R.A., 1984. Sedimentary pyrite formation: an update. Geochim. Cosmochim. Acta Hay, W.W., DeConto, R.M., Wold, C.N., Wilson, K.M., Voigt, S., Schulz, M., Rossby Wold, A.,
48, 605–615. Dullo, W.C., Ronov, A.B., Balukhovsky, A.B., Söding, E., 1999. Alternative global
Böttcher, M.E., Hetzel, S., Brumsack, H.-J., Schipper, A., 2006. Sulfur-iron-carbon Cretaceous paleogeography. In: Barrera, E., Johnson, C.C. (Eds.), Evolution of the
geochemistry in sediments of the Demerara Rise. In: Mosher, D.C., Erbacher, J., Cretaceous Ocean-climate System. Geol. Soc. Am. Spec., Paper 332, pp. 1–47.
Malone, M.J. (Eds.), Proc. ODP, Sci. Results, vol. 207, pp. 1–23. doi:10.2973/odp.proc. Helz, G.R., Miller, C.V., Charnock, J.M., Mosselmans, J.F.W., Pattrick, R.A.D., Garner, C.D.,
sr.2007.108.2006. Vaughan, D.J., 1996. Mechanism of molybdenum removal from the sea and its
Bornemann, A., Friedrich, O., Beckmann, B., Hofmann, P., Schouten, S., Sinninghe concentration in black shales: EXAFS evidence. Geochim. Cosmochim. Acta 60,
Damsté, J.S., Vogel, J., Wagner, T., Norris, R.D., 2008. Isotopic evidence for glaciation 3631–3642.
during the Cretaceous supergreenhouse. Science 319, 189–192. Hetzel, A., Brumsack, H.-J., Schnetger, B., Böttcher, M.E., 2006. Inorganic geochemical
Breit, G.N., Wanty, R.B., 1991. Vanadium accumulation in carbonaceous rocks: a review characterization of lithologic units recovered during ODP Leg 207 (Demerara Rise).
of geochemical controls during deposition and diagenesis. Chem. Geol. 91, 83–97. In: Mosher, D.C., Erbacher, J., Malone, M.J. (Eds.), Proc. ODP, Sci. Results, vol. 207.
Bruland, K.W., Donat, J.R., Hutchins, D.A., 1991. Interactive influences of bioactive trace Ocean Drilling Program, College Station, TX, pp. 1–37.
metals on biological production in oceanic waters. Limnol. Oceanogr. 36, Hetzel, A., Böttcher, M.E., Wortmann, U.G., Brumsack, H.-J., 2009. Paleo- redox
1555–1577. conditions during OAE 2 reflected in Demerara Rise sediment geochemistry
Brumsack, H.-J., 2006. The trace metal content of recent organic carbon-rich sediments: (ODP Leg 207). Palaeogeogr. Palaeoclimatol. Palaeoecol. 273, 302–328.
implications for Cretaceous black shale formation. Palaeogeogr. Palaeoclimatol. Huckriede, H., Meischner, D., 1996. Origin and environment of manganese-rich
Palaeoecol. 232, 344–361. sediments within black shale basins. Geochim. Cosmochim. Acta 60, 1399–1413.
Burdige, D.J., 1993. The biogeochemistry of manganese and iron reduction in marine Huerta-Diaz, M.G., Morse, J.W., 1992. Pyritization of trace metals in anoxic marine
sediments. Earth-Sci. Rev. 35, 249–284. sediments. Geochim. Cosmochim. Acta 56, 2681–2702.
Calvert, S.E., Pedersen, T.F., 1993. Geochemistry of recent oxic and anoxic marine Ingall, E.D., Jahnke, R.A.,1997. Influence of water-column anoxia on the elemental fractionation
sediments: implications for the geological record. Mar. Geol. 113, 67–88. of carbon and phosphorus during sediment diagenesis. Mar. Geol. 139, 219–229.
Chung, F.H., 1974. Quantitative interpretation of X-ray diffraction patterns, I. Matrix- Ingall, E.D., Bustin, R.M., Van Cappellen, P., 1993. Influence of water column anoxia on
flushing method of quantitative multicomponent analysis. J. Appl. Crystallogr. 7, the burial and preservation of carbon and phosphorus in marine shales. Geochim.
513–519. Cosmochim. Acta 57, 303–316.
Dean, W.E., Gardner, J.V., Piper, D.Z., 1997. Inorganic geochemical indicators of glacial– Jacobs, E., Emerson, S., 1982. Trace metal solubility in an anoxic fjord. Earth Planet. Sci.
interglacial changes in productivity and anoxia on the California continental Lett. 60, 237–252.
margin. Geochim. Cosmochim. Acta 61, 4507–4518. Junium, C.K., Arthur, M.A., 2007. Nitrogen cycling during the Cretaceous, Cenomanian–
Delaney, M.L., 1998. Phosphorus accumulation in marine sediments and the oceanic Turonian Oceanic Anoxic Event II. Geochem. Geophys. Geosyst. 8, Q03002.
phosphorus cycle. Glob. Biogeochem. Cycles 12, 563–572. doi:10.1029/2006GC001328.
Dellwig, O., Böttcher, M.E., Lipinski, M., Brumsack, H.-J., 2002. Trace metals in Holocene Karpoff, A.M., Destrigneville, C., Stille, P., 2007. Clinoptilolite as a new proxy of
coastal peats and their relation to pyrite formation (NW Germany). Chem. Geol. 182, enhanced biogenic silica productivity in lower Miocene carbonate sediments of
423–442. the Bahamas platform: isotopic and thermodynamic evidence. Chem. Geol. 245,
Dellwig, O., Beck, M., Lemke, A., Lunau, M., Kolditz, K., Schnetger, B., Brumsack, H.-J., 285–304.
2007. Non-conservative behaviour of molybdenum in coastal waters: coupling Kastner, M., 1980. Zeolites. In: Burns, R. (Ed.), Marine Minerals. Rev. Mineral., vol. 6.
geochemical, biological, and sedimentological processes. Geochim. Cosmochim. Mineralogical Society of America, pp. 111–122.
Acta 71, 2745–2761. Klinkhammer, G.P., Bender, M.L., 1980. The distribution of manganese in the Pacific
Erbacher, J., Mosher, D.C., Malone, M.J., 2004. Shipboard scientific party. Proc. ODP In. Ocean. Earth Planet. Sci. Lett. 46, 361–384.
Reports, vol. 207. doi:10.2973/odp.proc.ir.207.2004. Lewan, M.D., 1984. Factors controlling the proportionality of vanadium to nickel ratios
Erbacher, J., Friedrich, O., Wilson, P.A., Birch, H., Mutterlose, J., 2005. Stable organic in crude oils. Geochim. Cosmochim. Acta 48, 2231–2238.
carbon isotope stratigraphy across Oceanic Anoxic Event 2 of Demerara Rise, Lewan, M.D., Maynard, J.B., 1982. Factors controlling the enrichment of vanadium and
western tropical Atlantic. Geochem. Geophys. Geosyst. 6, Q06010. doi:10.1029/ nickel in the bitumen of organic sedimentary rocks. Geochim. Cosmochim. Acta 46,
2004GC000850. 2547–2560.
Feely, R.A., Massoth, G.J., Baker, E.T., Cowen, J.P., Lamb, M.F., Krogslund, K.A., 1990. The Lewis, B.L., Landing, W.M., 1991. The biogeochemistry of Mn and Fe in the Black Sea.
effect of hydrothermal processes on midwater phosphorus distributions in the Deep-Sea Res. 38, 773–803.
northeast Pacific. Earth Planet. Sci. Lett. 96, 305–318. Lewis, B.L., Luther, G.W., 2000. Processes controlling the distribution and cycling of
Filippelli, G.M., 2001. Carbon and phosphorus cycling in anoxic sediments of the Saanich manganese in the oxygen minimum zone of the Arabian Sea. Deep-Sea Res. II 47,
Inlet, British Columbia. Mar. Geol. 174, 307–321. 1541–1561.
C. März et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 286–301 301

Lovley, D.R., 1987. Organic matter remineralization with the reduction of ferric iron: a Schlanger, S.O., Jenkyns, H.C., 1976. Cretaceous oceanic anoxic events: causes and
review. Geomicrobiol. J. 5, 375–399. consequences. Geol. Mijnb. 55, 179–184.
Lunau, M., Lemke, A., Dellwig, O., Simon, M., 2006. Physical and biogeochemical controls Scholz, F., Neumann, T., 2007. Trace element diagenesis in pyrite-rich sediments of the
of microaggregate dynamics in a tidally affected coastal ecosystem. Limnol. Achterwasser lagoon, SW Baltic Sea. Mar. Chem. 107, 516–532.
Oceanogr. 51, 847–859. Schoonen, M.A.A., 2004. Mechanisms of sedimentary pyrite formation. In: Amend, J.P.,
März, C., Poulton, S.W., Beckmann, B., Küster, K., Wagner, T., Kasten, S., in press. Redox Edwards, K.J., Lyons, T.W. (Eds.), Sulfur Biogeochemistry–Past and Present. GSA
sensitivity of P cycling during marine black shale formation - dynamics of sulfidic Spec. Pap., vol. 379, pp. 117–134.
and anoxic, non-sulfidic bottom waters. Geochim. Cosmochim. Acta. Schulz, H.D., 2004. Quantification of Early Diagenesis: dissolved constituents in pore
Meyers, P.A., Bernasconi, S.M., Forster, A., 2006. Origins and accumulation of organic water and signals in the solid phase. In: Schulz, H.D., Zabel, M. (Eds.), Marine
matter in expanded Albian to Santonian black shale sequences on the Demerara Geochemistry. Springer, pp. 73–124.
Rise, South American margin. Org. Geochem. 37, 1816–1830. Shimmield, G.B., Price, N.B., 1986. The behaviour of molybdenum and manganese
Morford, J.L., Emerson, S.E., 1999. The geochemistry of redox sensitive trace metals in during early sediment diagenesis offshore Baja California, Mexico. Mar. Chem. 19,
sediments. Geochim. Cosmochim. Acta 63, 1735–1750. 261–280.
Morford, J.L., Emerson, S.R., Breckel, E.J., Kim, S.H., 2005. Diagenesis of oxyanions (V, U, Sinninghe Damsté, J.S., De Leeuw, J.W., 1990. Analysis, structure and geochemical
Re, and Mo) in pore waters and sediments from a continental margin. Geochim. significance of organically-bound sulphur in the geosphere: state of the art and
Cosmochim. Acta 69, 5021–5032. future research. Org. Geochem. 16, 1077–1101.
Morse, J.W., Luther III, G.W., 1999. Chemical influences on trace metal-sulfide Sinninghe Damsté, J.S., Kok, M.D., Köster, J., Schouten, S., 1998. Sulfurized carbohy-
interactions in anoxic sediments. Geochim. Cosmochim. Acta 63, 3373–3378. drates: an important sedimentary sink for organic carbon? Earth Planet. Sci. Lett.
Morse, J.W., Millero, F.J., Cornwell, J.C., Rickard, D., 1987. The chemistry of hydrogen 164, 7–13.
sulphide and iron sulphide systems in natural waters. Earth-Sci. Rev. 24, 1–42. Slomp, C.P., Van der Gaast, S.J., Van Raaphorst, W., 1996. Phosphorus binding by poorly
Mosher, D.C., Erbacher, J., Malone, M.J., 2007. Shipboard scientific party. Proc. ODP Sci. crystalline iron oxides in North sea sediments. Mar. Chem. 52, 55–73.
Results, vol. 207. doi:10.2973/odp.proc.sr.207.2007. Stone, A.T., Morgan, J.J., 1987. Reductive dissolution of metal oxides. In: Stumm, W. (Ed.),
Musavu-Moussavou, B., Danelian, T., 2006. The radiolarian response to oceanic anoxic Aquatic Surface Chemistry. Wiley, Chichester, pp. 221–254.
event 2 in the southern part of the Northern proto-Atlantic (Demerara Rise, ODP Leg Stumm, W., Sulzberger, B., 1992. The cycling of iron in natural environments:
207). Rev. Micropal. 49, 141–163. considerations based on laboratory studies of heterogeneous redox processes.
Nealson, K.H., Myers, C.R., 1992. Microbial reduction of manganese and iron: new Geochim. Cosmochim. Acta 56, 3233–3257.
approaches to carbon cycling. Appl. Environ. Microbiol. 58, 439–443. Tebo, B.M., 1991. Manganese(II) oxidation in the suboxic zone of the Black Sea. Deep-Sea
Nederbragt, A.J., Thurow, J., Pearce, R., 2007. Sediment composition and cyclicity in the Res. 38 (Suppl. 2), 883–905.
Mid-Cretaceous at Demerara Rise, ODP Leg 207. In: Mosher, D.C., Erbacher, J., Thurow, J., 1988. Diagenetic history of Cretaceous radiolarians, North Atlantic Ocean
Malone, M.J., et al. (Eds.), Proc. ODP Sci. Results, vol. 207, pp. 1–31. doi:10.2973/odp. (ODP Leg 103 and DSDP Holes 398D and 603B). In: Boillot, G., Winterer, E.L., et al.
proc.sr.207.103.2007. (Eds.), Proc. ODP Sci. Results, vol. 103, pp. 531–555.
Neumann, T., Heiser, U., Leosson, M.A., Kersten, M., 2002. Early diagenetic processes Torres, M.E., Brumsack, H.J., Bohrmann, G., Emeis, K.C., 1996. Barite fronts in continental
during Mn-carbonate formation: evidence from the isotopic composition of sediments: a new look at barium remobilization in the zone of sulfate reduction
authigenic Ca-rhodochrosites of the Baltic Sea. Geochim. Cosmochim. Acta 66, and formation of heavy barites in authigenic fronts. Chem. Geol. 127, 125–139.
867–879. Tribovillard, N., Riboulleau, A., Lyons, T., Baudin, F., 2004. Enhanced trapping of
Paytan, A., Griffith, E.M., 2007. Marine barite: recorder of variations in ocean export molybdenum by sulfurized marine organic matter of marine origin in Mesozoic
productivity. Deep-Sea Res. II 54, 687–705. limestones and shales. Chem. Geol. 213, 385–401.
Paytan, A., Kastner, M., Campbell, D., Thiemens, M.H., 2004. Seawater sulfur isotope Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox and
fluctuations in the Cretaceous. Science 304, 1663–1665. paleoproductivity proxies: an update. Chem. Geol. 232, 12–32.
Percy, D., Li, X., Taylor, G.T., Astor, Y., Scranton, M.I., in press. Controls on iron, manganese Tuit, C., Waterbury, J., Ravizza, G., 2004. Diel variation of molybdenum and iron in
and intermediate oxidation state sulfur compounds in the Cariaco Basin. Mar. marine diazotrophic cyanobacteria. Limnol. Oceanogr. 49, 978–990.
Chem. Turekian, K.K., Wedepohl, K.H., 1961. Distribution of the elements in some major units of
Pletsch, T., Erbacher, J., Holbourn, A.E.L., Kuhnt, W., Moullade, M., Oboh-Ikuenobede, the Earth's crust. Geol. Soc. Am. Bull. 72, 175–192.
F.E., Söding, E., Wagner, T., 2001. Cretaceous separation of Africa and South America: Van der Weijden, C., 2002. Pitfalls of normalization of marine geochemical data using a
the view from the West African margin (ODP Leg 159). J. S. Am. Earth Sci. 14, common divisor. Mar. Geol. 184, 167–187.
142–174. Vogt, C., Lauterjung, J., Fischer, R.X., 2002. Investigation of the clay fraction (b 2 µm) of
Poulton, S.W., Canfield, D.E., 2006. Co-diagenesis of iron and phosphorus in the clay mineral society reference clays. Clays Clay Miner. 50, 388–400.
hydrothermal sediments from the southern East Pacific rise: implications for the Wagner, T., 2002. Late Cretaceous to early Quaternary organic sedimentation in the
evaluation of paleoseawater phosphate concentrations. Geochim. Cosmochim. Acta eastern equatorial Atlantic. Palaeogeogr. Palaeoclimatol. Palaeoecol. 179, 113–147.
70, 5883–5898. Wagner, T., Pletsch, T., 1999. Tectono-sedimentary controls on Cretaceous black shale
Poulton, S.W., Krom, M.D., Raiswell, R., 2004. A revised scheme for the reactivity of iron deposition along the opening Equatorial Atlantic Gateway (ODP Leg 159). In:
(oxyhyr)oxide minerals towards dissolved sulfide. Geochim. Cosmochim. Acta 68, Cameron, N., Bate, R., Clure, V. (Eds.), The Oil and Gas Habitats of the South Atlantic.
3703–3715. Geol. Soc. London. Spec. Pub., vol. 153, pp. 241–265.
Redfield, A.C., 1958. The biological control of chemical factors in the environment. Am. Wagner, T., Sinninghe Damsté, J.S., Hofmann, P., Beckmann, B., 2004. Euxinia and
Sci. 64, 205–221. primary production in Late Cretaceous eastern equatorial Atlantic surface waters
Reitz, A., Wille, M., Nägler, T., De Lange, G.J., 2007. Atypical Mo isotope signatures in fostered orbitally driven formation of marine black shales. Paleoceanography 19,
eastern Mediterranean sediments. Chem. Geol. 245, 1–8. PA3009. doi:10.1029/2003PA000898.
Rosenthal, Y., Lam, P., Byle, E.A., Thomson, J., 1995. Authigenic cadmium enrichments in Wehrli, B., Stumm, W., 1989. Vanadyl in natural waters: adsorption, and hydrolysis
suboxic sediments: precipitation and postdepositional mobility. Earth Planet. Sci. promote oxygenation. Geochim. Cosmochim. Acta 53, 69–77.
Lett. 132, 99–111. Yakushev, E.V., Pollehne, F., Jost, G., Kuznetsov, I., Schneider, B., Umlauf, L., 2007. Analysis
Ruttenberg, K.C., 1992. Development of a sequential extraction method for different of the water column oxic/anoxic interface in the Black and Baltic seas with a
forms of phosphorus in marine sediments. Limnol. Oceanogr. 37, 1460–1482. numerical model. Mar. Chem. 107, 388–410.
Schenau, S.J., De Lange, G.J., 2000. A novel chemical extraction method to quantify fish Yao, W., Millero, F.J., 1996. Oxidation of hydrogen sulfide by hydrous Fe(III) oxides in
debris in marine sediments. Limnol. Oceanogr. 45, 963–971. seawater. Mar. Chem. 52, 1–16.
Schenau, S.J., Slomp, C.P., De Lange, G.J., 2000. Phosphogenesis and active phosphorite Zheng, Y., Anderson, R.F., van Geen, A., Kuwabara, J.S., 2000. Authigenic molybdenum
formation in sediments from the Arabian Sea oxygen minimum zone. Mar. Geol. formation in marine sediments: a linkage to pore water sulfide in the Santa Barbara
169, 1–20. Basin. Geochim. Cosmochim. Acta 64, 4165–4178.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Paleo-redox conditions during OAE 2 reflected in Demerara Rise sediment


geochemistry (ODP Leg 207)
Almut Hetzel a,⁎, Michael E. Böttcher b,c, Ulrich G. Wortmann d, Hans-Jürgen Brumsack a
a
Institute for Chemistry and Biology of the Marine Environment (ICBM), Carl von Ossietzky University, P.O. Box 2503, D-26111 Oldenburg, Germany
b
Max Planck Institute for Marine Microbiology, Celsiusstrasse 1, D-28359 Bremen, Germany
c
Leibniz Institute for Baltic Sea Research Warnemünde, Seestrasse 15, D-18119 Warnemünde, Germany
d
Department of Geology, University of Toronto, 22 Russel street, Ontario, Canada M5S 3B1

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

Article history: Cretaceous black shales deposited under severe oxygen-depletion are carriers of proxy signals for paleoenviron-
Received 1 December 2007 mental conditions. Using high-resolution patterns of iron and sulfur speciation, stable sulfur isotope
Received in revised form 8 August 2008 discrimination, and trace element enrichment from black shale sequences of Sites 1258 and 1260 we identified
Accepted 3 November 2008
alterations of the depositional environment during the Cenomanian/Turonian boundary Event (OAE 2) in the
southern North Atlantic (ODP Leg 207, Demerara Rise).
Keywords:
Demerara Rise
Changes in redox-conditions are suggested by high ratios of reactive to total iron which indicate that pyrite was
Ocean Drilling Program formed both in the water column and within the sediment. This corresponds to euxinic paleoenvironmental
Leg 207 conditions with at least temporarily free dissolved sulfide in the water column, a situation similar to the modern
Trace metals deep Black Sea. In addition, besides fixation of sulfide as iron sulfide, organic matter acted as an important sulfur
Oceanic Anoxic Event 2 trap during early diagenesis. Stable sulfur isotope fractionation went through a minimum within the OAE 2 interval
CTBE indicating enhanced sulfur isotope discrimination during highest burial of organic matter (OM) potentially due to
Paleoenvironment lower burial efficiency of reduced sulfur and/or a higher contributions from the oxidative part of the sulfur cycle
Pyrite
(e.g., in the water column or the surface sediments). Elevated Fe/Al and Co/Al values within the Cenomanian/
Sulfur isotopes
Turonian interval confirm euxinic conditions but, at the same time, require a zone where reducing but non-sulfidic
conditions prevail, allowing reductive Fe and Co mobilization in oxygen-depleted nearshore sediments. The
existence of an expanded oxygen-minimum-zone (OMZ) is demonstrated by extremely low Mn/Al ratios.
A change in the trace metal (TM) inventory of seawater is postulated from a decline in seawater derived TM
enrichment. Because hypoxic or even euxinic environments form an important sink for TM, the enlargement of
euxinic depositional areas at the global onset of black shale deposition during OAE 2 have likely led to a drawdown
of the seawater TM reservoir.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction characterized by a global organic carbon burial episode leading to a


positive shift in δ13C values of organic carbon and carbonate (e.g.,
In the mid-Cretaceous, several distinct periods of organic-rich black Schlanger et al., 1987; Arthur et al., 1988; Gale et al.,1993; Erbacher et al.,
shale deposition appear. The enhanced burial of organic carbon in 2005). As the carbon cycle is tightly coupled to the sulfur cycle, pertur-
marine sediments during these so called Oceanic Anoxic Events (OAEs; bations in the global carbon cycle lead to changes in the seawater sulfur
Schlanger and Jenkyns, 1976) is thought to arise either from enhanced isotopic composition (e.g., Strauss, 1999; Paytan et al., 2004; Wortmann
bio-productivity or from intensified preservation of organic matter and Chernyavsky, 2007).
during anoxic conditions (e.g., Arthur et al., 1987, 1988; Schlanger et al., During Ocean Drilling Program (ODP) Leg 207, relatively expanded,
1987). The Cenomanian–Turonian Boundary Event (CTBE = OAE 2; ca. shallowly buried Cretaceous sediments were recovered from Demer-
93.5 Ma) is one of the best studied global Oceanic Anoxic Events. It is ara Rise off Suriname, South America. CTBE black shales on Demerara
Rise form part of a thick black shale succession encompassing the
Albian to Santonian (Shipboard Scientific Party, 2004). This may be the
result of both the paleogeographic setting of Demerara Rise (lacking
significant ventilation of bottom-waters prior to the opening of the
⁎ Corresponding author.
E-mail addresses: almut.hetzel@gmx.de (A. Hetzel),
equatorial Atlantic gateway, which may have taken place in the
michael.boettcher@io-warnemuende.de (M.E. Böttcher), brumsack@icbm.de Campanian; see Friedrich and Erbacher, 2006) and enhanced nutrient
(H.-J. Brumsack). influx. The latter may have been evoked by terrestrial runoff and/or

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.11.005
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 303

upwelling due to the proximal position of Demerara Rise to the South 2. Material and methods
American landmass.
Cretaceous black shales deposited under severe oxygen-depletion 2.1. Site description
are carriers of proxy signals for paleoenvironmental conditions like
varying levels of bottom water dysoxia and/or enhanced surface water During ODP Leg 207, sediment cores were drilled on the Demerara
productivity. Changes in water column redox conditions lead to re- Rise in the tropical North Atlantic (Fig. 1). The rise stretches ca. 380 km
sponses in the coupled biogeochemical sulfur–carbon–metal cycles and along the coast off Suriname and reaches a width of ca. 220 km from the
associated sedimentary signal formation. Therefore, distribution pat- shelf break to the northeastern escarpment, where water depths increase
terns of iron and sulfur speciation, sulfur isotope partitioning, and sharply from 1000 to more than 4500 m. While most of the plateau lies in
enrichments of redox-sensitive and sulfide forming trace metals provide shallow water (700 m), the northwest margin forms a gentle ramp
important paleoenvironmental information on dynamics of biogeo- reaching water depths of 3000 to 4000 m. The five drill sites (Sites 1257–
chemical element cycles and corresponding water column redox- 1261) constitute a depth transect with water depths ranging from
conditions during black shale deposition. Excess iron contribution to 1900 to 3200 m located within a tropical oxygen minimum zone that
selected Cretaceous Leg 207 black shales due to pyrite formation in a caused deposition of laminated organic-rich sediments of latest Albian to
euxinic water column was shown by iron and sulfur speciation analyses, earliest Campanian age. Upper Cenomanian to Lower Turonian sedi-
and enhanced organic matter sulfurization was shown to be caused by a ments on Demerara Rise are mainly expressed by distinctly laminated
limitation of reactive iron to form sulfide minerals (Böttcher et al., 2006). black shales with well-preserved fish debris and phosphatic nodules.
High enrichments of redox-sensitive elements in organic carbon- Within this black shale sequence light-colored, laminated foraminiferal
rich sediments have been related to anoxic bottom waters. Under packstones and wackestones occur (Shipboard Scientific Party, 2004). A
reducing conditions these metals may either be precipitated as detailed sedimentological description is given by Erbacher et al. (2004).
sulfides, co-precipitated with iron sulfides or bound to organic matter Linear sedimentation rates (LSRs) were estimated by the Shipboard
(Brumsack, 1980; Jacobs et al., 1985, 1987; Brumsack, 1989; Breit and Scientific Party (2004) from age-depth plots by fitting curves to available
Wanty, 1991; Hatch and Leventhal, 1992; Calvert and Pedersen, 1993; biostratigraphic and paleomagnetic data over certain depth intervals.
Piper, 1994; Nijenhuis et al., 1998). Based on the specific trace metal Poor preservation or even absence of microfossils and the ambiguous
patterns of coastal upwelling areas and euxinic settings Brumsack interpretation of paleomagnetic data limit the bio- and magnetostrati-
(2006) discussed the enrichment of trace metals in Cretaceous black graphic age assignment in most of the black shale sequences (Cenomanian–
shales attempting to ascertain whether enhanced bio-productivity or Santonian). Therefore, LSRs calculated for these intervals (0.3–0.5 cm/ka
widespread stagnation triggered black shale formation. at Sites 1257, 1258, and 1259 and slightly higher values of ~0.85 cm/ka at
In the present study, we report on a high-resolution geochemical Sites 1260 and 1261) should be considered as imprecise estimates.
tracer record for black shale sequences of ODP Leg 207 Sites 1258 and 1260 Stable isotope stratigraphy provided by Erbacher et al. (2005)
using distribution patterns of trace element enrichment, iron and sulfur allows a detailed correlation of the investigated sites through the OAE
speciation, and stable sulfur isotope discrimination to identify changes in 2 interval. At all studied sites, a distinctive positive carbon isotope
redox-conditions of the depositional environment during OAE 2. excursion (~6.5‰) of the OAE 2 was identified (Fig. 2). The same authors

Fig. 1. Geographic position of Sites 1257–1261 at Demerara Rise (ODP Leg 207). (Online Map Creation www.aquarius.geomar.de).
304 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Fig. 2. Depth-profiles for TOC contents for the whole sediment column investigated (open circles) (Hetzel et al., 2006) and for Cenomanian/Turonian black shales at Demerara Rise,
ODP Sites 1258 and 1260. Grey: OAE 2; δ13Corg-curve after Erbacher et al. (2005).

were able to correlate a number of short-termed δ13Corg peaks and at Site 1260). (3) An interval “C” with δ13Corg values rising to −23 and
troughs within and above the excursion interval between the −21‰. (4) The final maximum value “D”. (5) A short positive peak “E”
investigated sites. The events of this record include: (1) The onset of occurs during the general decrease in isotope values (not present at Site
the excursion labeled “A”. (2) A short-termed minimum “B” (not present 1258). In addition, Erbacher et al. (2005) also correlated two peaks (“F”
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 305

and “G”) located above the OAE 2 between the sites. Assuming a duration Table 2
of ~400 ka for the interval between “A” (onset of excursion) and “D” (end Sampling Site 1258 and 1260

of plateau) Erbacher et al. (2005) calculated average sedimentation rates Site Sampled Number of Time of Sampling Sampling
between ~0.25 cm/ka (Site 1260) and ~1 cm/ka (Site 1258) for the OAE 2. interval samples deposition resolution resolution
There is no more detailed information available about sedimentation (mcd) (Ma) (cm) (ka)

rates for the intervals above and below OAE 2 or variations within this 1258 415.0–448.5 845 ~ 3.35 ~ 1/~ 10 ~ 1/~ 10
1260 423.5–432.0 219 ~ 3.40 ~5 ~ 20
interval.
In this study, two ODP sites along the Demerara Rise paleodepth- mcd = meters composite depth.
transect (Sites 1258 and 1260, see Table 1) were chosen to reconstruct
paleoenvironmental changes during the late Cenomanian to earliest ELTRA® CS-500 IR-analyzer. Total inorganic carbon (TIC) was determined
Turonian black shale deposition interval. Depths are expressed in meters coulometrically by a UIC® CM 5012 CO2 coulometer coupled to a CM 5130
composite depth (mcd), following the shipboard splice between the acidification module. Total organic carbon (TOC) was calculated as the
different holes drilled at each site (Shipboard Scientific Party, 2004). difference between TC and TIC. Procedures and accuracy of the methods
were checked with in-house reference materials (Prakash Babu et al.,
2.1.1. ODP Site 1258 1999; see Appendix A). Different sedimentary sulfur fractions, chromium-
Site 1258 is located in a water-depth of 3192.2 m below modern sea reducible sulfur (SP, essentially pyrite), organic matter (essentially
level (mbsl) on the gently dipping western slope (~2°) of Demerara Rise. It kerogen-bound organic sulfur, SORG) and acid volatile sulfur (SAVS) were
is the deepest location in the SE–NW paleoceanographic depth transect separated quantitatively from powdered freeze-dried samples, as
across Demerara Rise (Shipboard Scientific Party, 2004). The 4 m-thick described by Böttcher et al. (2006). Chromium reducible sulfur, essentially
OAE 2, as defined by the pronounced δ13Corg excursion, is located at a pyrite sulfur, SP, was extracted for 2 h via hot acidic Cr(II)Cl2 (Zhabina and
depth of 426–422 mcd (Erbacher et al., 2005). Based on δ13Corg measure- Volkov, 1978; Canfield et al., 1986; Fossing and Jørgensen, 1989). Liberated
ments, Erbacher et al. (2005) described a hiatus in the uppermost part of the H2S was precipitated quantitatively in Zn acetate traps and measured
OAE. The sediments of the investigated section comprise finely-laminated spectrophotometrically (Cline, 1969). The organic sulfur fraction, SORG,
dark shales with occasional beds of phosphatic nodules, stringers of was calculated as the difference of total sulfur and the sum of chromium-
very dark homogeneous shales, and rare concretionary limestone nodules. reducible sulfur (Böttcher et al., 2006). SAVS was extracted from selected
Thirty-three and a half meters (415–448.5 mcd, see Table 2) of upper samples from Site 1258 via anaerobic distillation with 6 M HCl (1 h). Since
Cenomanian to lowermost Turonian laminated black shale sediments FeS is not expected in the black shale samples to survive the diagene-
were investigated from a spliced interval of Holes 1258A, 1258B, and tic pyritization and lab-based freeze-drying process, the SAVS fraction is
1258C. 845 samples were taken at 1 cm resolution for the OAE 2 interval assumed to correspond to acid-volatile metal monosulfides. ZnS, for
and at ~10 cm in underlying sediments. Assuming a sedimentation rate of instance, that might have been formed in a euxinic water column, has
~1 cm/ka the studied interval covers a time span of ~3.35 Ma of deposition. been found in C/T samples in the present study via SEM-EDX, and has also
been reported in Cretaceous black shales previously (Brumsack, 1980).
2.1.2. ODP Site 1260 ZnS would have survived burial and later sample handling.
Site 1260, in a water-depth of 2548.8 mbsl, is positioned on the gently The amount of pyrite iron (FeP) was calculated from the SP content
dipping (~1°) northwest-facing slope of Demerara Rise, at an inter- assuming ideal stoichiometry. This fraction essentially represents the total
mediate depth in the SE–NW paleoceanographic depth transect. The highly reactive iron fraction (Raiswell and Canfield, 1998; Poulton and
1.4 m thick OAE 2 is located between 426.4 and 425.0 mcd, being thus Raiswell, 2002). To confirm this approach, iron-oxide bound iron was
much thinner than at Site 1258. Lithologically, the interval is very similar separately determined in selected black shale samples from Sites 1258 and
to Site 1258, although phosphatic nodules are not as common as at the 1260 and presented by Böttcher et al. (2006). As described in detail by
deeper site (Erbacher et al., 2005). About 8.5 m of upper Cenomanian to these authors, iron(oxyhydr)oxide phases were extracted from sediment
lowermost Turonian laminated black shale sediments were sampled samples using a buffered solution of Na dithionite and extracted iron was
every ~5 cm (219 samples in total, see Table 2) from a spliced interval of determined spectrophotometrically. It was found that this fraction is only
Holes 1260A and 1260B. Sedimentation rates of the OAE 2 interval are present in very small amounts and that pyrite iron amounts to 96–99% of
estimated to be on the order of approximately 0.25 cm/ka, thus leading the total highly reactive iron pool. For samples from Site 1258 34S/32S ratios
to the same duration of ~3.4 Ma of deposition for the studied sequence. of the SP bound sulfur were determined by means of C-irmMS using a
Eurovector 3000 elemental analyzer (EA) coupled to a Thermo Finnigan
2.2. Analytical methods MAT253 and are given in the δ-notation versus the international V-CDT
standard. Calibration of the mass spectrometer was carried using interna-
Sediments splits were freeze-dried, ground and homogenized in an tional IAEA and NBS reference materials. Sulfur isotope results for Site 1260
agate ball mill. For X-ray fluorescence (XRF) analysis (Philips® PW 2400 will be presented in a separate communication (Böttcher et al., in prep.).
X-ray spectrometer), 600 mg of sample powder were mixed with Finally, selected non-ground carbon-coated (BAL-TEC sputter-coating
3600 mg of a 1:1 mixture of dilithiumtetraborate (Li2B4O7) and lithium- device SCD 005) freeze-dried sediment samples were investigated by
metaborate (LiBO2), or with 100% dilithiumtetraborate for carbonate- means of energy-dispersive X-ray analysis on an Oxford Link ISIS 300
rich samples, preoxidized at 500 °C with NH4NO3 (p.a.) and fused to glass EDX-System with Pentafet S ATW Si-detector installed on a Zeiss DSM
beads. Total sulfur (ST) and total carbon (TC) were analyzed using an 940 scanning electron microscope (SEM).

Table 1
3. Results and discussion
Drilling locations Site 1258 and 1260, ODP Leg 207
3.1. Preliminary remarks
Site Hole Latitude Longitude Water depth OAE 2 interval SR
(mbsl) (mcd) (cm/ka)
To account for dilution effects by varying carbonate or organic matter
1258 A 9° 26.000′N 54° 43.999′W 3192.2 425.95–422.18 ~1
1258 B 9° 26.000′N 54° 43.982′W 3192.2 contents trace element contents were normalized to Al (TE/Al ratio). In
1258 C 9° 26.000′N 54° 43.966′W 3192.2 cases of very low Al contents (see Figs. 3 and 4), normalization may lead
1260 A 9° 15.984′N 54° 32.633′W 2548.8 426.34–424.94 ~ 0.25 to exaggerated peaks in TE/Al ratios, which has to be kept in mind when
1260 B 9° 15.931′N 54° 32.652′W 2548.8 interpreting TE/Al profiles (van der Weijden, 2002). Some elements, like
mbsl = meters below sea level; mcd = meters composite depth; SR = sedimentation rate. Al, Ti, K, Rb and Zr, are only present in the detrital component and are not
306 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

influenced either by biogenic or diagenetic processes. Significant To better compare TOC and ST contents for Demerara C/T black shales
variations in TE/Al ratios of these elements therefore reveal changes in with other C/T sections we calculated TOCcf and ST cf for a carbonate-free
source area composition. While most of the investigated sediment sediment. Profiles are given in Figs. 3 and 4 beside profiles of absolute
intervals show TE/Al ratios similar to average shale (AS; Wedepohl, 1971, concentrations. For Site 1258 most of the TOC values vary between ~0.5
1991), several sediment layers of one centimeter to several decimeters and ~15%. Adjusting for dilution by carbonate resulting TOCcf values are
thickness were identified by characteristic TE/Al ratios. At Site 1260 a quite higher (~10–~25%). Some intervals of half a meter thickness are
35 cm thick interval (425.61–425.96 mcd) within the OAE 2 is characterized by TOC contents N20%. These peaks are distributed right
characterized by low K/Al (0.07, AS = 0.34), high Ti/Al ratios (0.102, below, within and above OAE 2. For Site 1260 the TOC contents vary
AS = 0.053) and is depleted in Si (Si/Al 2.58, AS = 3.11; see Appendix B). between ~0.5 and ~12% and on a carbonate free basis TOCcf values fall
From macroscopic observation (Appendix B-b), this interval most likely between ~15 and ~22%. Highest TOC contents (N15%) are found within
is a diagenetically altered ash bed. At Site 1258, three sediment layers OAE 2.
within OAE 2 (423.51–423.53 mcd, 423.87–424.04 mcd, and 425.22– Average contents below, within and above OAE 2 (Table 3) show
425.50 mcd) correspond to this layer, identified by similar disturbances higher TOC and lower TIC contents for Site 1258 in comparison to Site
in TE/Al profiles. At both sites abrupt changes in carbonate, TOC and TS 1260. For both sites, average TOC values are in the same range (TOC:
contents from extremely high concentrations to near zero values can be 6.9–10.1%; TOCcf: 17.6–19.3%) below and above the OAE 2 and slightly
identified. Light δ13C values of the carbonate in the discussed interval at higher during the OAE 2 with 17.8% (TOCcf 22.4%) at Site 1258 and 12.1%
Site 1260 (Friedrich, pers. communication) point towards the involve- (TOCcf 22.8%) at Site 1260. The same is true for ST values with 1.5–2.4%
ment of microbiological processes like anaerobic oxidation of methane (ST cf 3.8–4.7%) below and above and 5.0% (ST cf 6.3%) within the OAE 2
(AOM) (e.g., Galimov, 2006). The production of bicarbonate by AOM at Site 1258 and 3.4% (ST cf 6.5%) within the OAE 2 at Site 1260.
could result in the precipitation of authigenic carbonates, and thus
increase the degree of carbonate diagenesis implying recrystalization of 3.3. Iron
carbonates from foraminifera and nanofossils (Erbacher et al., 2004).
With the geochemical methods applied (bulk element concentrations) Figs. 3 and 4 show profiles of total iron and Al-normalized iron
the origin of the material can not be identified. To avoid misinterpreta- contents. Only for some carbonate-rich layers Fe contents are below the
tion of changes in element distribution patterns, samples from these quantification limit of 0.33% (e.g. ~424 mcd at Site 1260) and FeT/Al
intervals (7 samples from Site 1260 and 27 samples from Site 1258) were peaks may be exaggerated (see Section 3.1.) due to inaccuracies of
excluded from following discussion regarding changes of redox- measurements of Fe and Al close to or below the quantification limit.
conditions and trace metal availability. As the bulk geochemistry of Besides these peaks, the FeT/Al profiles show an almost constant value
sites is similar, analytical results of discussed intervals are given as close to AS. Assuming severe oxygen-deficiency, as evidenced by
average values listed in Appendix B-c. sediment lamination for the black shale sequence of Demerara Rise,
the almost constant average shale-like values of FeT/Al imply anoxic but
3.2. Bulk parameters not euxinic (no free H2S in the water column) conditions during
deposition. FeT/Al ratio of the sediment is neither decreased via
Fig. 2 shows the depth profiles of total organic carbon (TOC) for the mobilization of iron(hydr)oxides under suboxic conditions nor
whole sediment column (data from Hetzel et al., 2006) and the upper increased via precipitation of dissolved Fe species as sulfide in the
Cenomanian to lower Turonian black shale sequences (C/T back overlying water column under euxinic conditions. During OAE 2 FeT/Al
shales), which are in the focus of this study. mean values increase to 0.76 (Site 1258, maximum 1.72) and 0.80 (Site
Sediments above these sequences are characterized by low TOC 1260, maximum 1.89), owing to an increase in total FeT contents. This
contents (b0.4%). Below these intervals TOC contents decrease from indicates an additional Fe-source during deposition of this interval. In
~ 10% to ~1% with increasing depth. The C/T black shales show varying analogy to the modern Black Sea reductive Fe mobilization in near-shore
TOC values: lowest contents (almost 0%) in carbonate-rich layers and areas or enhanced fluvial input are likely causes for the observed Fe-
highest contents (30.2% at Site 1258 and 23.1% at Site 1260) within the enrichment (e.g., Lyons and Severmann, 2006). In the case of the modern
OAE 2 interval, which is identified by correlation with the δ13Corg Black Sea FeT/Al-ratios are increasing with water depth (e.g. Wijsman
curve from Erbacher et al. (2005) and marked in grey. et al., 2001) because syngenetic iron sulfides form an increasing part of
Figs. 3 and 4 show the bulk chemistry of the C/T black shales: as bulk sedimentary iron. This model requires truly euxinic conditions,
mentioned above, highly variable TIC contents (b0.2–11.8%) lead to besides the existence of a zone where reducing but non-sulfidic
dilution effects, which are shown by the inverse distribution pattern conditions prevail, allowing Fe transport from shallow to deep sites
for TIC and Al, representing terrigenous detritus. The TIC profiles (e.g., Lyons and Severmann, 2006). The increase of the ratio of pyrite iron
display the light–dark cycles described by the Shipboard Scientific to total iron, FeP/FeT, from a mean value of 0.44 below and above to about
Party. Whereas the profile of Site 1260 suggests a regular cyclicity 0.64 (Table 4) within OAE 2 indicates the addition of syngenetic water
possible modulated by orbital forcing, carbonate contents of Site 1258 column-derived pyrite (Raiswell and Canfield, 1998; Poulton and
show a less regular distribution pattern despite higher resolution (see Raiswell, 2002; Böttcher et al., 2006, and references therein). Together
Table 2). For both sites, time series analysis of our high-resolution TIC with previous measurements on the detailed reactive iron speciation of
data does not evidence orbital cyclicity. Leg 207 sediments (Böttcher et al., 2006), these findings are confirmed
Nederbragt et al. (2007) analyzed the cyclicity during the Mid- by the Al-normalized iron contents, as discussed above. Fig. 5 presents
Cretaceous at Demerara Rise. Due to the irregular spacing of the the analytical data of Sites 1258 and 1260 in a ternary Fex–ST–TOC plot
carbonate-rich and organic-rich cycles (lack of precise age control, (Brumsack, 1988; Dean and Arthur, 1989; Brumsack et al., 1995). The
variable sedimentation rates and/or degree of compaction and minor represented data within the OAE all plot below the pyrite-saturation line
hiatuses), the authors concluded instead of establishing a continuous indicating, that under iron-limited conditions, some (excess) sulfide
cyclostratigraphy for the entire organic-rich unit to perform time series and/or sulfur intermediates were able to react with organic matter to
analysis of sediment color data in selected intervals in combination with form organic sulfur compounds.
thin section analysis of representative lithologies. They found that cyclic
variation in lithology at Demerara Rise is inferred to represent eccentricity 3.4. Sulfur and sulfur isotopes
and precession cycles with a weak obliquity component. For analysis of
the cyclicity during OAE 2 the authors examined two cores from Site 1258 Dissimilatory sulfate reduction leads to the formation of hydrogen
and 1260. Both, precession and eccentricity cycles were found. sulfide that may transform reactive iron to iron sulfides (essentially
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 pp. 307–308

Fig. 3. Depth-profiles for TOC, TIC, Al, ST, SORG, SP, SAVS and FeT for Cenomanian/Turonian black shales at Site 1258. TOCcf and ST cf are calculated for carbonate-free sediment. δ34S of
pyrite, FeT/Al ratios as well as FeP/FeT ratios are also shown. Grey: OAE 2; grey solid line: element/Al ratio of to ‘average shale’ (AS) (Wedepohl, 1971, 1991); dashed line: quantification
limit (see Appendix A).
pp. 309–310 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Fig. 4. Depth-profiles for TOC, TIC, Al, ST, SORG, SP and FeT for Cenomanian/Turonian black shales at Site 1260. TOCcf and ST cf are calculated for carbonate-free sediment. FeT/Al ratios as
well as FeP/FeT ratios are also shown. Grey: OAE 2; grey solid line: element/Al ratio of ‘average shale’ (AS) (Wedepohl, 1971, 1991); dashed line: quantification limit (see Appendix A).
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 311

Table 3
Comparison of average element contents and element/Al ratios Cenomanian/Turonian black shales of Demerara Rise and ‘average shale’

1258 below OAE 2 1258 within OAE 2 1258 above OAE 2 1260 below OAE 2 1260 within OAE 2 1260 above OAE 2 Average shale
(n = 347) (n = 197) (n = 274) (n = 129) (n = 26) (n = 57) (Wedepohl, 1971, 1991)
ST % 1.71 4.97 2.35 1.45 3.43 1.69 0.24
TIC % 6.20 2.46 5.00 7.59 5.76 6.34 0.35
TOC % 7.83 17.82 10.13 6.89 12.12 7.54
Si % 11.61 14.66 11.43 6.64 8.06 9.14 27.53
Ti % 0.08 0.14 0.12 0.08 0.10 0.10 0.46
Al % 1.68 2.90 2.57 1.47 2.17 2.18 8.84
FeT % 0.88 2.17 1.19 0.76 1.65 0.97 4.80
Mg % 0.47 0.46 0.57 0.47 0.47 0.60 1.60
Ca % 21.37 9.44 18.27 26.04 20.20 22.32 1.57
Na % 0.68 1.16 0.99 0.74 0.86 0.82 1.19
K % 0.45 0.89 0.74 0.43 0.65 0.69 2.99
P % 0.25 0.15 0.54 0.29 0.22 0.52 0.07
As ppm 13 25 21 18 25 16 10
Ba ppm 260 542 612 343 746 642 580
Co ppm 3 13 5 3 9 4 19
Cr ppm 121 70 149 109 82 145 90
Cu ppm 58 68 70 53 68 63 45
Mn ppm 7 50 27 0 14 7 850
Mo ppm 80 40 95 62 33 38 1
Ni ppm 126 169 149 108 144 92 68
Rb ppm 21 30 30 19 22 26 140
Sr ppm 593 653 716 923 952 792 300
U ppm 12 17 18 16 18 12 4
V ppm 1058 370 1173 1386 554 899 130
Y ppm 12 29 23 13 23 19 41
Zn ppm 578 138 968 879 382 701 95
Zr ppm 28 38 36 27 33 34 160
ST cf % 3.83 6.26 4.36 4.04 6.47 4.70 0.25
TOCcf % 17.64 22.40 18.30 18.74 22.80 19.30
ST/Al 1.05 1.75 1.05 1.06 1.67 1.54 0.03
TOC/Al 4.85 6.32 4.40 4.94 5.93 5.98
Si/Al 7.79 5.81 4.76 4.49 3.72 4.24 3.11
Ti/Al 0.05 0.05 0.05 0.05 0.05 0.05 0.05
FeT/Al 0.55 0.76 0.50 0.58 0.80 0.74 0.54
Mg/Al 0.57 0.17 0.25 0.55 0.25 1.39 0.18
Ca/Al 42.97 4.36 11.69 43.01 13.87 121.32 0.18
Na/Al 0.43 0.41 0.42 0.57 0.42 0.58 0.13
K/Al 0.28 0.31 0.29 0.32 0.31 0.40 0.34
P/Al 0.13 0.05 0.20 0.17 0.12 0.23 0.01
As/Al ×10− 4 9.3 8.8 8.4 14.3 12.5 13.6 1.1
Ba/Al ×10− 4 190.8 188.1 239.0 271.8 345.2 531.6 65.6
Co/Al ×10− 4 2.1 4.4 2.3 2.2 4.4 3.2 2.1
Cr/Al ×10− 4 68.8 24.0 50.7 75.1 37.5 57.7 10.2
Cu/Al ×10− 4 36.0 24.0 27.4 38.9 30.5 45.9 5.1
Mn/Al ×10− 4 9.2 18.4 23.2 0.1 5.9 36.2 96.2
Mo/Al ×10− 4 50.0 14.3 40.4 49.4 16.7 20.3 0.1
Ni/Al ×10− 4 77.8 59.5 62.0 85.0 73.0 47.6 7.7
Rb/Al ×10−4 11.5 10.2 11.3 12.1 10.7 10.6 15.8
Sr/Al ×10−4 584.4 241.5 349.0 1019.4 541.5 1425.7 33.9
U/Al ×10− 4 9.7 6.2 8.5 14.5 10.8 17.9 0.4
V/Al ×10− 4 654.8 127.0 461.7 1010.1 294.5 439.8 14.7
Y/Al ×10− 4 8.9 10.1 10.2 10.6 13.1 17.0 4.6
Zn/Al ×10− 4 316.0 48.1 355.9 581.4 154.0 251.1 10.7
Zr/Al ×10− 4 20.7 13.5 14.9 22.4 17.1 31.8 18.1

Italics: average values below quantification limit (see Appendix A).

pyrite, FeS2; including framboidal pyrite Fig. 6) and, under iron limited during OAE 2 at both sites (Figs. 3 and 4; see also Table 4). Below OAE 2 SP
conditions with organic matter to form organic sulfur compounds (e.g. values are ~0.4% (ca. 25 rel% of ST) and SORG ~1.2% (respectively ca. 75 rel%
Werne et al., 2004). The profiles of SP and SORG show elevated values of ST). At Site 1258, SP contents rise to ~1.7% (maximum N2.5%) and SORG
to ~4.4% (maximum N7.7%) within the OAE 2 interval, thus the proportion
Table 4 of pyrite sulfur increases to ca. 28 rel% of total sulfur. Although absolute
Average pyrite sulfur (SP), pyrite iron (FeP), acid volatile sulphur (SAVS), calculated
organic sulfur (Sorg) and ratio of pyrite (reactive) iron to total iron (see Table 3)
values are not as high as at Site 1258, the pyrite sulfur fraction at Site 1260
reaches even ca. 33 rel% with 1.1% SP (maximum N3.1%) and 2.3% SORG
1258 1258 1258 1260 1260 1260 (maximum N6.6%) during OAE 2. Above OAE 2 the proportional
below within above below within above
composition of different sulfur species nearly equals the composition
OAE 2 OAE 2 OAE 2 OAE 2 OAE 2 OAE 2
(n = 19) (n = 18) (n = 12) (n = 16) (n = 22) (n = 13) below with ~0.5% SP and ~1.4% SORG (SP:SORG =26:74 rel%) at both sites.
SP % 0.42 1.68 0.60 0.38 1.14 0.52 The relationship of SP with TOC data is presented in Fig. 7a and compared
FeP % 0.36 1.46 0.52 0.33 0.99 0.45 to the relationship proposed for “normal marine sediments” (Berner and
SAVS ppm 22.4 (n=17) 0.5 (n=13) 12.2 (n=11) Raiswell, 1983). Only a few data points coincide with the relation found
Sorg % 1.25 4.39 1.69 1.14 2.31 1.41 for clastic sediments below an oxic water column (dashed line in Fig. 7a).
FeP/FeT 0.41 0.67 0.44 0.43 0.60 0.46
A number of data points plot above the regression line, indicating an
312 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Fig. 5. Ternary plot: degree of pyritization of Cenomanian/Turonian black shales at Demerara Rise in the Fex–TOC–ST · 2 system (relative weight ratios). Reactive Fex is calculated with
Fex = FeT − 0.25 · Al. Data point for pyrite is also shown.

excess of sulfur that may be due to euxinic depositional conditions. At variations observed in the sedimentary pyrite pool are much more
highest TOC contents, however, most data show a relative excess of pronounced than those observed in the global seawater sulfate at this
organic matter (OM). The indicated iron-limitation during black shale time (Paytan et al., 2004). This indicates that sulfur isotope discrimi-
deposition has also been found for the modern Black Sea sapropel and nation was controlled by changes in sedimentary conditions and was
Mediterranean sapropels (e.g. Lyons and Berner, 1992; Passier et al., higher during times of enhanced burial of OM. This can be caused by a
1999b; Böttcher et al., 2006). SP in the investigated black shale samples lower burial efficiency of reduced sulfur and/or higher contributions
makes up between 30% and 100% of ST, with a decrease of the relative from the oxidative part of the sulfur cycle (in the water column or in
amount of SP with increasing OM content. This indicates the importance surface sediments). A similar trend and magnitude of sulfur isotope
of the balance between organic matter and the syngenetic metal flux to discrimination was found in the C/T sediments of different sites in the
the surface sediments in controlling the sedimentary sulfur speciation. In Tarfaya basin (e.g. Kolonic et al., 2002; Böttcher et al., unpublished data).
addition to fixation of sulfide by the reaction with iron, organic matter For instance for the pyrite fraction Böttcher et al. (unpublished data)
acted as the second important sulfur trap during early diagenesis. From obtained ranges at Site S13 (−18.9 to +2.5‰, average −10.2 ±7.8‰; n =9),
molecular analysis of the preservation pathways of sedimentary organic at S57 (−28.8 to −8.8‰, average −17.6 ±5.1‰; n =29), and S75 (−20.6 to
carbon in a euxinic environment Hebting et al. (2006) proposed that −8.7‰, average −15.7± 3.2‰; n =17). This similarity suggests a common
sulfides produced by bacterial sulfate reduction play a major role for the control of the overall sedimentary sulfur isotope signal for the CTBE in the
reductive alteration and thus preservation of organic carbon. From the southern North Atlantic independent of paleo-water depths. The
nearly linear variation of organic sulfur and organic carbon contents stratigraphic trend of the sulfur isotope ratios shows that pyrite isotope
(Fig. 7b), essentially constant atomic SORG/TOC ratios are obtained. data go through a minimum similar to previous observations at Tarfaya
Quantitatively, the samples with TOC contents exceeding ~2 wt.% have as (Kolonic et al., 2002; Böttcher et al., unpublished data). We relate the
much as 10 atom% organic sulfur. Most of the atomic S/C ratios fall in the most negative values to an increased contribution of syngenetic pyrite,
range of 0.04 to 0.06, which is within the range reported for sapropels sulfur derived from the oxidative part of the biogeochemical sulfur cycle,
from the Mediterranean and Black Sea (Passier et al., 1999b; Böttcher and mostly changes in other factors (e.g., the quality of organic matter)
et al., 2006). The almost linear relationship between TOC and SORG may be influencing overall sulfur isotope discrimination. Most lighter sulfur
caused by a constant relative proportion of organic matter fractions isotope data are found during OAE 2 compared to sediments outside OAE
available for diagenetic sulfurization. 2 (Fig. 8), and these are associated with highest contents of pyrite-iron.
The downcore variation of the stable sulfur isotope composition of This is in agreement with findings of Gauthier (1987) on Cretaceous black
pyrite at Site 1258 displays a range in δ34S values between −28.2 and shales from the Western Interior Seaway. Taking the sulfur isotopic
+0.2‰ (average −15.4 ± 8.3‰; n= 41) during OAE 2, with an inverse composition for seawater sulfate during the Cenomanian/Turonian to be
relationship to the δ13C values (Fig. 3, Table 5). The sulfur isotope around +15‰ (Strauss, 1999; Paytan et al., 2004), and assuming that
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 313

Fig. 6. EDX-Scan and SEM of a pyrite framboid from Site 1258B-45-4 138–139 cm, 422.98 mcd (within OAE 2).

pyrite mainly formed under conditions that were open with respect and 426 ppm, with a mean value of ~50 ppm and a standard deviation of
to dissolved sulfate (Hartmann and Nielsen, 1969), e.g., in the water the same order (49 ppm). Above OAE 2 (421.63–421.03 mcd) a clear peak
column or close to the sediment–water interface, an overall sulfur in absolute Mn concentrations with a maximum value of 217 ppm can be
isotope fractionation of 15 to 42‰ is estimated for Site 1258. This is recognized. This peak can be correlated to a peak in TIC. Since Mn
within the range found in studies with pure cultures of sulfate-reducing concentrations often are below the quantification limit, Mn/Al ratios
bacteria (e.g. Brunner and Bernasconi, 2005), but is smaller than values have to be considered with caution. A maximum Mn/Al ratio of 466× 10− 4
found in C/T sediments from the northern North Atlantic (Böttcher and at 421.63–421.03 mcd represents a significant Mn enrichment in
Kuypers, unpublished data), in the Pliocene TOC-rich sapropels of the comparison to the Mn/Al ratio of 96 ×10− 4 for average shale (Wedepohl,
deep Mediterranean (Passier et al., 1999a; Böttcher et al., 2003), in the 1971, 1991). The same is true for certain intervals within OAE 2. For
modern euxinic Black Sea (Neretin et al., 2004; Böttcher et al., 2004), and samples close to the onset and termination of OAE 2, the enrichments in
in sediments from the Great Australian Bight (Wortmann et al., 2001). Mn are less pronounced.
Higher contributions from the oxidative part of the sulfur cycle (microbial At Site 1260 Mn concentrations in all samples are below the
disproportionation of sulfur intermediates) and/or lower cellular sulfate quantification limit. Only a few samples attain the detection limit of
reduction rates or specific bacterial communities may be responsible for 39 ppm: One interval above OAE 2 (423.99–424.04 mcd) reaching Mn
the development of the different characteristic sulfur isotope signals. values up to 77 ppm can again be correlated to a carbonate peak.
Within OAE 2 Mn concentrations vary between 0 and 46 ppm.
3.5. Redox-sensitive and sulfide forming trace metals Extremely low Mn/Al ratios in C/T black shales of Demerara Rise
indicate that Mn was either reduced within the water column before
3.5.1. Manganese sedimentation or upon early diagenesis in an environment open to
The C/T black shales of Demerara Rise are characterized by very low dissolved Mn(II) loss as provided by an expanded oxygen-minimum-
Mn-concentrations (Fig. 9). For most of the samples the Mn contents are zone (OMZ), like the modern upwelling zones off Peru (Böning et al.,
below the XRF quantification limit of 78 ppm. At Site 1258, for sediments 2004) or in the Gulf of California (Brumsack, 1989). Thurow et al.
below 427.7 mcd Mn values are below the detection limit, which is (1992) found a similar mobilization of Mn at the Northwest Australian
defined as half the quantification limit. Right below, within and above margin during OAE 2. The authors described Mn-poor sediments
OAE 2, Mn contents are higher and reach the detection limit of 39 ppm. within the OMZ and Mn-rich sediments below the OMZ, indicating
Within OAE 2, Mn values show a high variability between detection limit oxic deep waters during the C/T interval at this location.
314 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Fig. 8. δ34S of pyrite versus pyrite iron (FeP) for Cenomanian/Turonian black shales at
Site 1258.

ponsible for Mn enrichment in Black Sea sediments. Pre-existing carbo-


nates may act as an efficient trap for diagenetically mobilized Mn(II)
(Boyle, 1983; Gingele and Kasten, 1994; Böttcher, 1997).
At Demerara Rise elevated Mn/Al ratios are only seen in distinct
horizons with high carbonate contents below (Site 1258) and above OAE
2 (Site 1258 and Site 1260). We therefore relate these Mn/Al peaks in C/T
black shales to carbonate diagenesis. There is no correlation to FeT/Al
ratios that would point to simultaneous mobilization from suboxic
sediments and precipitation under sulfidic conditions as described by
Lyons and Severmann (2006).
The higher Mn contents at Site 1258 might originate from the
position in greater paleo-water depths. Assuming the existence of an
OMZ, Site 1258 might be located below the OMZ where during times
of weaker anoxia, Mn(IV)-bearing phases reached the sediment and
later were transformed into mixed Ca–Mn-carbonates.
Within OAE 2, Mn values show a high variability as described
above. As all OAE 2 samples with higher Mn/Al ratios are located near
the presumed “ash-layers”, the import of Mn(IV)-bearing phases could
also be attributed to the different detritus composition (average Mn
contents: 332 ppm compared to 17 ppm mean value for all C/T black
shales of this study, see Appendix B-c). If Mn was released to the pore
space after burial, Mn concentrations in the pore water within these
“ash-layers” and adjacent intervals might have been higher. Because
TIC contents are not elevated (1.7–8.1%, mean value 5.6%), precipita-
tion of Mn-carbonate is questionable in these intervals.
Besides increased availability of dissolved Mn, pore volume may
Fig. 7. (a) Pyrite sulfur (SP) versus TOC and (b) organic sulfur (SORG) versus TOC for
influence the potential for precipitation of Mn-bearing minerals during
Cenomanian/Turonian black shales at Site 1258 and Site 1260 (this study) and black shale
samples from Sites 1257, 1259 and 1261 (Böttcher et al., 2006). Dashed line marks the
diagenesis. The carbonate-rich layers with elevated Mn contents are
relationship derived for normal marine sediments as defined by Berner and Raiswell (1983). visually affected by carbonate diagenesis, e.g. precipitation of calcite
(Erbacher et al., 2004). Increased pore volume, either through the
A possibility for the observed Mn-enrichment might be the fixation occurrence of foraminiferal packstones or of “ash-like” layers, may
of dissolved Mn(II) as mixed Mn–Ca-carbonates under conditions where have favored authigenic formation of Mn-bearing minerals.
Mn was enriched. Brumsack (2006) suggests that overgrowths of early
diagenetic Mn-carbonate on pre-existing carbonate tests might be res- 3.5.2. Cobalt
Under oxygen depleted conditions like in an OMZ, Mn–and Fe-oxi/
hydroxide-associated Co would be mobilized and transported away. In
Table 5
Average δ34S of pyrite sulfur (SP) for Cenomanian/Turonian black shales of Site 1258 Gulf of California upwelling sediments (Brumsack, 1989) or the Peruvian
margin (Böning et al., 2004), Co is depleted similar to Mn. In contrast to
1258 below OAE 2 1258 within OAE 2 1258 above OAE 2
Mn, which forms stable sulfides only under very special conditions, Co
(n=16) (n=12) (n=13)
precipitates as CoS when free hydrogen sulfide is present (Heggie and
δ34SP ‰ −8.4 −24.4 −15.6
Lewis, 1984; Gendron et al., 1986).
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 315

Fig. 9. Depth-profiles for Mn, Mn/Al, Co and Co/Al for Cenomanian/Turonian black shales at Site 1258 and Site 1260. Grey: OAE 2; grey solid line: element/Al ratio of ‘average shale’
(AS) (Wedepohl, 1971, 1991); dashed line: quantification limit (see Appendix A).

The profiles of Co concentration (Fig. 9) at Site 1258 and Site 1260 requires an additional Co-source during deposition, we assume a trans-
show that nearly all sediments below and above OAE 2 are characterized port mechanism analogous to Fe from near-shore areas, where reducing
by very low Co concentrations (often below the quantification limit of but non-sulfidic conditions mobilize additional Co, which is then fixed
9 ppm). Co/Al ratios for Demerara Rise C/T black shales below and above under truly euxinic conditions further offshore.
OAE 2 are similar to the Co/Al ratio of average shale (2.1 × 10− 4;
Wedepohl, 1971, 1991). Within OAE 2, Co/Al ratios increase to a mean 3.5.3. Zinc
value of 4.4 × 10− 4 at both sites with maximum values of 9.4 × 10− 4 at Site Cretaceous black shales are known to be highly enriched in Zn.
1258 and 10.3 × 10− 4 at Site 1260. Brumsack (2006) gives a Zn/Al ratio 459 × 10− 4 as a C/T mean value
Thus, the Co distribution is quite similar to the Fe distribution: Under (AS = 10.7 ×10− 4; Wedepohl, 1971, 1991). Hetzel et al. (2006), found
reducing conditions the constant average shale-like value of Co/Al rather 370 ×10− 4 as a mean value of the Cretaceous black shales of Demerara
indicates fixation of dissolved Co as sulfide than preservation of parti- Rise (all sites). The high-resolution study of the C/T interval presented
culate Co-bearing (hydr)oxide phases. As the Co increase during OAE 2 here reveals considerable variability in Zn/Al ratios (Fig. 10). The depth
316 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

profile of Site 1258 shows Zn/Al ratios in the same order of magnitude as For anoxic basins, Emerson and Huested (1991) showed that the
described by Hetzel et al. (2006) below and above OAE 2 (mean values of concentrations of Mo and V in the water column are usually lower than
316.0 × 10− 4 and 355.9 ×10− 4, respectively; see Table 3). The maxima in oxic seawater due to their uptake into highly anoxic sediments. As no
below and above OAE 2 correspond to maxima in the SAVS profile, indi- systematic relationship between the deep-water Mo or V concentrations
cating the presence of metal sulfides other than pyrite. At Site 1260, the and H2S content was found, the authors concluded that concentrations
Zn/Al ratios are slightly higher below OAE 2 (mean value 581.4 ×10− 4), of Mo and V are rather controlled by their flux to sediments and water
while above OAE 2 ratios are lower (mean value 251.1 × 10− 4). At both sites renewal from outside the basins than by changes in water column
a distinct decrease in Zn/Al ratios during OAE 2 can be observed (mean anoxia. Algeo and Lyons (2006) analyzed Mo–TOC relationships in
values 48.1 × 10− 4 at Site 1258 and 154.0 ×10− 4 at Site 1260). sediments from anoxic environments characterized by different degrees
Addressing the distribution patterns of Co and Zn discussed above of hydrographic restriction. In silled anoxic basins, decreasing Mo/TOC
we found two different enrichments patterns for these trace metals, ratios with increasing degree of restriction imply that sedimentary Mo
both known to be enriched in C/T black shales (Brumsack, 2006) by concentrations were controlled by Mo availability from the overlying
forming stable sulfides. Whereas C/T black shales from Demerara Rise water column, and thus by resupply of Mo via deepwater renewal. Algeo
are enriched in Co during OAE 2, but not before and thereafter, the and Lyons (2006) termed this drawdown of Mo concentrations in deep
enrichment in Zn is present throughout the studied interval and water during stagnant intervals the “basin reservoir effect”.
displays a rapid decrease in the degree of enrichment within OAE 2. Fig. 11 shows Mo/TOC profiles as well as X–Y-plots of Mo versus TOC.
Mo/TOC values vary between ~5 × 10− 4 and ~20× 10− 4 for sediments
3.5.4. Molybdenum and vanadium deposited before OAE 2. Similar to Mo/Al, Mo/TOC ratios decrease rapidly
Overall, element/Al ratios of both Mo and V show clear enrichments at the onset of OAE 2, staying low during OAE 2, before a slight increase is
in C/T black shales of Demerara Rise (mean Mo/Al = 28.2 × 10− 4, visible after OAE 2. The slopes of regression lines for X–Y-plots of Mo
AS = 0.15 × 10− 4, mean V/Al = 443.8 × 10− 4, AS = 14.7 × 10− 4; Wedepohl, versus TOC give similar low Mo/TOC ratios for samples within the OAE 2
1971, 1991; Fig. 10). Mo and V are relatively unreactive in oxic seawater, interval at both sites (2.3× 10− 4 at Site 1258 and 2.4× 10− 4 at Site 1260).
but are known to be concentrated in sediments overlain by anoxic These values are even lower than the value of Mo/TOC= 4.5 ± 1 × 10− 4
waters (Brumsack and Gieskes, 1983; Brumsack, 1986). reported by Algeo and Lyons (2006) for the Black Sea. Regression lines for
Vanadium in oxic seawater should be present as V(V). Under samples deposited before OAE 2 give Mo/TOC values of 10.7× 10− 4 at Site
moderately reducing conditions, V(IV) forms vanadyl ions (VO2−) that 1258 and 8.8× 10− 4 at Site 1260. The increase in Mo/TOC ratios after OAE
may be removed to the sediment by surface adsorption processes or 2 is less pronounced at Site 1260 (Mo/TOC = 5.2× 10− 4) than at Site 1258
by formation of organometallic ligands (Emerson and Huested, 1991; (Mo/TOC = 9.4 × 10− 4).
Morford and Emerson, 1999). Under more reducing conditions, the As V burial under anoxic conditions is also known to be linked to OM
presence of free H2S released by bacterial sulfate reduction causes V to (Breit and Wanty, 1991; Emerson and Huested, 1991; Morford and
be further reduced to V(III), which can be taken up by geoporphyrins Emerson, 1999), we analyzed V/TOC ratios analogues to Mo and Zn/ST
or be precipitated as the solid oxide V2O3 or hydroxide V(OH)3 phase ratios as Zn is known to form stable sulfides if H2S is present (Jacobs et al.,
(Breit and Wanty, 1991; Wanty and Goldhaber, 1992). 1985) (Fig. 11). Both, V and Zn, show the distinct decrease relative to TOC
The stable oxidation state of Mo in oxic seawater is Mo(VI). Reduction and ST during OAE 2 pointing to decoupling of OM burial, sulfidization
of Mo(VI) to Mo(IV) and authigenic enrichment in sediments occur under and TM availability during OAE 2.
euxinic conditions (Crusius et al., 1996) possibly via diffusion across the Fig.12 shows the enrichment factors relative to average shales for FeT,
sediment–water interface (Emerson and Huested, 1991). Helz et al. Co, Mo, V and Zn for Sites 1258 and 1260, and for C/T mean values
(1996) suggested a threshold concentration for H2S, above which Mo is (Brumsack, 2006). The enrichment factor for FeT is ~1 at Site 1258 below
transformed to particle-reactive thiomolybdates that are scavenged by and above the CTBE and at Site 1260 below the CTBE, indicating Fe/Al
forming bonds with metal-rich particles, sulfur-rich OM and pyrite. ratios similar to average shale. The observed enrichment in sediments of
Tribovillard et al. (2006) addressed the reconstruction of the redox Site 1260 above the CTBE may be an artefact owing to normalization of
environment of sediments overlain by oxygen-depleted waters via the values close to the Fe and Al detection limit in carbonate-rich sediments
combined use of different trace metals. While V is reduced and can (see Fig. 4, ~424 mcd). As described above, the distribution patterns of Co
accumulate under denitrifying conditions, Zn and Mo are enriched are similar to those of FeT.
mainly under sulfate-reducing conditions. Thus, in the case of V Zn, Mo and V show similar enrichment patterns, with highest values
enrichment without Mo enrichment, the authors postulate suboxic/ below, intermediate values above, and lowest values within OAE 2. At
anoxic depositional conditions without free H2S, whereas sediments Site 1258, Mo, V and Zn are enriched to similar degrees below and above
exhibiting concurrent enrichments in V and Mo reflect euxinic OAE 2. The decline of enrichment during OAE 2 is more pronounced than
conditions at the sediment–water interface or in the water column for Site 1260. Overall, the degrees of enrichment of Mo, V and Zn are
(Algeo and Maynard, 2004; Tribovillard et al., 2004). As the C/T black similar to other C/T sediments listed by Brumsack (2006). Only the
shales of Demerara Rise are clearly enriched in both, Mo and V, this sediments deposited during OAE 2 at Site 1258 are characterized by less
simplified distinctive feature would hint to euxinic conditions for the pronounced Mo, V and Zn enrichments relative to other C/T black shales.
depositional environment. In summary, we suggest two important mechanisms affecting FeT and
From positive correlations of Mo with sulfurized OM in ancient trace metal distribution patterns in C/T black shales at Demerara Rise:
marine black shales, Tribovillard et al. (2004) emphasized Mo trapping Locally, Mn depletion hints to reducing conditions an open-marine
by sulfur-rich OM. No correlation of Mo with pyrite accumulation could environment, e.g. an OMZ. FeT and Co enrichments during OAE 2 indicate
be found. Instead, the reduced availability of reactive Fe may favor Mo an expansion of this OMZ. As their seawater concentrations are generally
accumulation, as significant OM sulfurization is only possible when low, at least part of the sedimentary FeT and Co enrichments during OAE 2
reactive iron is limited. As shown above, we find iron-limitation in must be due to mobilization from oxygen-depleted nearshore-sediments.
most of the studied intervals and an enhanced OM sulfurization during Whereas Fe and Co are conveyed along with Mn in suboxic waters, they
OAE 2. Following the arguments of Tribovillard et al. (2004), we would will be precipitated as sulfides when hydrogen sulfide is present in the
expect an increase in Mo enrichment during OAE 2. Instead, a clear water column.
decrease is visible, implying that other factors influenced Mo Regarding the source for TMs enriched in TOC-rich sediments, Mo
accumulation as well. Probably, the concentration of dissolved sulfide and V are highly concentrated in seawater, and they may have become
in the water column that affected aqueous Mo speciation may have enriched in the sediment via diffusion from the water column. As
influenced the efficiency of Mo fixation (Neubert et al., in press). extreme Zn enrichments in Cretaceous black shales (compared to
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 pp. 317–318

Fig. 10. Depth-profiles for Mo, Mo/Al, V, V/Al, Zn and Zn/Al for Cenomanian/Turonian black shales of Site 1258 and Site 1260. Grey: OAE 2; grey solid line: element/Al ratio of ‘average
shale’ (AS) (Wedepohl, 1971, 1991); dashed line: quantification limit (see Appendix A).
pp. 319–320 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Fig. 11. (a) Depth-profiles for Mo/TOC, V/TOC and Zn/ST for Cenomanian/Turonian black shales at Site 1258 and Site 1260. Grey: OAE 2. (b) X–Y-Plots for Mo versus TOC, V versus TOC
and Zn versus ST. Slopes of regression lines (solid for Site 1258, dashed line for Site 1260) are given in units of 10− 4.
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 321

Fig. 12. Mean enrichment factors (EF) of FeT, Co, Mo, V and Zn relative to ‘average shale’ (AS) (Wedepohl, 1971, 1991) for Cenomanian/Turonian black shales at Demerara Rise.
EF = element/Al(sample)/element/Al(AS). (a) Site 1258, (b) Site 1260. Dashed lines: EF for C/T mean values (Brumsack, 2006).

recent TOC-rich sediments) require an additional Zn source, a higher 3.6. Productivity proxies
seawater concentration of Zn due to hydrothermal input during the
Cretaceous seems likely (Arthur et al., 1988; Brumsack, 2006). Under Fig. 13 show the enrichments factors for some elements, used for
such conditions seawater might again be the dominating source for Zn approximation of paleoproductivity.
enrichments. Thus, on a larger if not global scale, we emphasize a Phosphorus belongs to the essential nutrient elements controlling
decline in seawater TM availability during OAE 2, revealed by the marine primary productivity (e.g. Broecker and Peng, 1982). High
decrease in Mo, V and Zn enrichment. concentrations are found in recent upwelling sediments pointing
Taking seawater as the dominant source for Mo, V and Zn, the towards enhanced nutrient supply and resulting high bioproductivity
depletion in TM/Al may reflect the global onset of black shale (e.g., Böning et al., 2004). Incorporated into organic material (e.g. in
deposition during OAE 2 and the rapid drawdown of the seawater marine phytoplankton; Redfield, 1958) P is deposited in the sediment,
TM reservoir by the spreading of euxinic depositional areas. In the where it is preferentially remineralized relative to TOC and liberated to
modern oceans, only 0.2% of the seafloor are hypoxic (Helly and Levin, the pore water and maybe forming authigenic phosphatic precipitates.
2004), while during the Cretaceous, particularly during OAE 2, such Under oxygen-depleted conditions, when potential phosphate scaven-
conditions were much more wide-spread. Kuypers et al. (2002) gers such as Fe-(hydr)oxides are reduced and not available, organic P
pointed out that the proto-North Atlantic Ocean was one of the main release in sediments is even higher. Therefore, in TOC-rich sediments
sites of carbon burial during OAE 2. Paleogeographic reconstructions deposited under anoxic conditions the TOC/P ratio should be increased
show that the proto-North Atlantic Ocean was a restricted area with (Ingall et al., 1993 and references therein).
major seawater supply coming from the Tethys only. The black shales Nederbragt et al. (2004) concluded from measurements of total
deposited in the Tethys and in the proto-North Atlantic Ocean during phosphorus and organic carbon in C/T sediments from the Tarfaya Basin,
OAE 2 (e.g. Kuhnt et al., 1990) definitely formed a significant TM sink. Morocco, that the main underlying mechanism that allowed and
We therefore assume that the decline in TM enrichment seen at sustained enhanced carbon burial during the mid-Cretaceous was a
Demerara Rise indicates a drawdown of TMs from seawater, similar to perturbation of the oceanic phosphorus cycle. More effective regeneration
the “basin reservoir effect” described by Algeo and Lyons (2006). of organic phosphorus under anoxic conditions can lead to an increase in
322 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

dissolved oceanic phosphate, which in turn stimulates surface-water degradation but also anaerobic oxidation of methane (AOM) above the
productivity (Ingall et al., 1993; Bjerrum and Canfield, 2002; Mort et al., black shales of Demerara Rise influences sulfate availability and therefore
2007). The Shipboard Scientific Party (2004) describes the presence of P- the remobilization of biogenic barium. These authors further showed that
rich concretions in black shales of Demerara Rise. Fig. 13 shows a clear temporal dynamics of degradation processes caused various shifts of the
enrichment of P relative to average shale in the sections studied. This barite precipitation zone during burial, thus inhibiting the formation of an
points towards a high primary productivity comparable to coastal authigenic barite front or causing the dissolution of earlier formed fronts.
upwelling areas. The minor enrichment during OAE 2 hints to a decreasing Fig. 13 shows an enrichment of Ba similar to the C/T mean values
nutrient supply or to partial loss of P during at least periodically truly from Brumsack (2006). This enrichment indicates elevated primary
anoxic conditions in comparison to the situation before and thereafter. productivity during deposition. But a large fraction of former barite
The non-lithologenic excess barium has been interpreted as a may have been remobilized and formed diagenetic barites above the
paleoproxy for bio-productivity (Schmitz, 1987; Dymond et al., 1992; black shales. For this reason the use of Ba as a paleoproxy on a
Paytan et al., 1996). These biogenic barites (BaSO4) (Bishop, 1988; quantitative level (Dymond et al., 1992) for Cretaceous settings in such
Bertram and Cowen, 1997; Paytan et al., 2002; Bernstein and Byrne, an environment is highly questionable.
2004) are only stable under seawater sulfate concentrations (Church and
Wohlgemuth, 1972). Due to the microbial sulfate reduction in TOC-rich 4. Paleoenvironmental implications and comparison with
sediments, barite may be dissolving, leading to the mobilization of Ba other paleoproxies
(Brumsack and Gieskes, 1983; McManus et al., 1998; Eagle et al., 2003).
Authigenic barite may precipitate at the top of the sulfate-depletion The paleoconditions for the depositional environment based on
zone forming diagenetic barite fronts within or above TOC-rich strata geochemical arguments are summarized in Fig. 14. From analysis of
(Torres et al., 1996; Bréhéret and Brumsack, 2000). TOC–Fe–S systematics and TM distribution patterns we can conclude
Hetzel et al. (2006) found the highest Ba/Al above the Cretaceous black that a depositional environment with an expanded oxygen-minimum-
shales, while within the black shales Ba/Al ratios are still high despite of zone (OMZ) similar to coastal upwelling areas existed before the onset of
the present absence of sulfate in the pore waters (Erbacher et al., 2004). OAE 2 (Fig. 14a). Enrichments of TOC, P and Ba hint to enhanced
Arndt et al. (2006) showed in a transport-reaction model that not only OM bioproductivity. The depletion in Mn and enrichment in redox-sensitive

Fig. 13. Mean enrichment factors (EF) of Si, P and Ba relative to ‘average shale’ (AS) (Wedepohl, 1971, 1991) for Cenomanian/Turonian black shales at Demerara Rise. EF = element/Al
(sample)/element/Al(AS). (a) Site 1258, (b) Site 1260. Dashed lines: EF for C/T mean values (Brumsack, 2006).
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 323

TM indicate at least suboxic conditions. During OAE 2, an increase in the shale’-like terrigenous background for C/T-black shales of Demerara
FeP/FeT ratio points to the rapid development of truly euxinic conditions Rise the enrichment in Si indicates the presence of additional Si other
with free hydrogen-sulfide in the water column (Fig. 14b). This is than present in the clay component. Owing to the absence of heavy
confirmed by a rise in FeT/Al and Co/Al ratios. An expansion of the OMZ, mineral related elements Hetzel et al. (2006) could exclude the
possible due to intensified paleoproductivity and/or sea level rise, leads presence of excess quartz in Cretaceous black shales of Demerara Rise
to mobilization of Fe and Co from nearshore sediments. Thus, and instead assumed a biogenous origin for the excess silica. We
enrichment in FeT and Co in the sediments is possible, when hydrogen suppose that the enrichment in Si shown in Fig. 13 represents Si
sulfide is present in the water column for sulfide precipitation. derived from radiolarian tests. The general pattern with slight
Musavu-Moussavou and Danelian (2006) analyzed the radiolarian enrichment of Si below and above OAE 2 fits well with the findings
abundances across OAE 2 at Site 1258 and Site 1261. The presence of of Musavu-Moussavou and Danelian (2006).
radiolarians at Site 1258 and to a lesser degree at the shallower Site Hardas and Mutterlose (2007) analyzed the calcareous nannofossil
1261 below OAE 2 indicates, at least periodically, oxygenated surface assemblages of Site 1258 and 1260. They found highest absolute abun-
waters. During OAE 2, the absence of radiolarians at Site 1261 and only dances below OAE 2 at both sites. During OAE 2 the absolute abundance
rare occurrence at Site 1258 hint at the intensification of euxinic decreases and also the species assemblages change. From the now
conditions, while above OAE 2 radiolarians are present again, this time dominating species the authors conclude a higher bio-productivity.
with higher abundances at Site 1261 than at Site 1258. In marine Instead of cooler temperatures, as suggested by Forster et al. (2007) at
sediments Si derives either from the terrigenous detritus as alumino- Site 1260 within OAE 2, Hardas and Mutterlose (2007) postulate a slight
silicates, quartz or from a biogenous opal from siliceous plankton increase in sea surface temperatures. They suggest a possible break-
(diatoms or radiolarians; Brumsack, 1989). Assuming an ‘average down of water column stratification and thus a shallowing of the

Fig. 14. Simplified scheme of paleoenvironmental conditions as deduced from geochemical data before/after (a) and during (b) OAE 2.
324 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

nutricline. This fits into a picture of a euxinic water column and the overall sulfur cycle, similar to results observed previously in the Tarfaya
decline in total abundance of nannofossils. Basin.
Regarding the redox conditions at the sediment–water interface
According to the assumed high bioproductivity and the depletion in
Friedrich et al. (2006) suggest a strengthening of the OMZ during OAE
Mn, the C/T black shales of Demerara Rise are compatible with a
2, based on the benthic foraminifera assemblage at Sites 1258–1261.
situation encountered in modern coastal upwelling areas within an
Below OAE 2 the authors suggest anoxic to sometimes slightly
expanded OMZ. During OAE 2 a more euxinic situation was established
dysoxic bottom-water conditions at the shallower sites, whereas for the
with the presence of free hydrogen sulfide in the water column.
deepest Site 1258 more oxygenated but still dysoxic bottom waters were
proposed, possibly due its position below the OMZ. During OAE 2, the
Acknowledgements
total absence of benthic foraminifera hints to anoxic conditions, but
short-term re-populations may indicate bottom-water oxygenation. The We would like to thank the crew and scientific party of ODP Leg 207
authors find these events in the lower third of OAE 2 at all sites studied. for their kind support. This research used samples and/or data provided
Forster et al. (2007) correlate these occurrences with a cooling event at by the Ocean Drilling Program (ODP). ODP is sponsored by the U.S.
Site 1260. Benthic foraminiferal assemblages indicate continued oxygen- National Science Foundation (NSF) and participating countries under
deficient bottom-water conditions following OAE 2 with the exception of management of Joint Oceanographic Institutions (JOI), Inc. This contribu-
one short-term increase in oxygenation immediately above the OAE 2. tion has benefited from the thorough reviews by Tim Lyons and an
After ~0.5 Ma. more oxygenated bottom-water masses are postulated. anonymous reviewer. AH is grateful to Astrid Forster and Christian März
From geochemical data we suppose euxinic conditions during OAE 2. for valuable discussions, Knut Bernhardt for performing time series
The presence of benthic foraminifera indicates that the euxinic conditions analysis and Imke Notholt for EDX assistance. MEB wishes to thank Andrea
may not have been permanent. Evidence for the short-term complete re- Schipper for technical assistance and E. Clapton, N. Futado, C. Hynde, and
oxidation of the bottom water is not given by the geochemical data, J. Stone for acoustical inspiration during work on the manuscript. This
although peaks in Mn/Al several decimeters above the repopulation study was funded by Deutsche Forschungsgemeinschaft (IODP-SPP grants
events of benthic foraminifera at Site 1258 may indicate a shallowing of BR 775/16+17; BO 1584/2) and by Max Planck Society, Germany.
the OMZ thus introducing Mn to the sediment.
Addressing the same questions by different paleo-proxies may lead
to different results under certain circumstances. If sedimentation rates Appendix A
are low, and the amount of sample material required for certain analyses Quantification limit, precision and accuracy of analyzed elements
comprises a time-span in which several changes in depositional Element Method Quantification limit Precision SD (1σ) (rel%) Accuracy (rel%)
environment may have occurred, the chemical result would be an
ST IR-analyzer 0.56% 3.0 99.9
integrated signal due to homogenization of the sample. In contrast, TC IR-analyzer 1.22% 1.6 101.2
analyzing fossils might give only a snapshot of paleoconditions. TIC Coulometry 0.19% 2.3 99.8
Contrasting results for the same intervals may therefore hint to unstable Si XRF 1.94% 1.3 99.1–101.0
Ti XRF 0.03% 1.2 98.1–101.0
conditions. Forster et al. (2007) suggest such unstable conditions during
Al XRF 0.77% 1.6 98.7–100.6
OAE 2 at Site 1260. From our point of view a closer look is necessary, if FeT XRF 0.33% 1.2 98.7–100.6
and to what extent the presence of the “ash-like” layers, which were Mg XRF 0.26% 1.9 99.3–101.0
omitted here, may influence directly or by diagenetic effects results Ca XRF 0.80% 1.6 99.4–99.7
obtained in close vicinity to these layers. Na XRF 0.21% 5.7 90.1–100.6
K XRF 0.20% 2.6 95.5–100.4
P XRF 0.04% 6.5 98.6–108.8
5. Summary and conclusions As XRF 4 ppm 7.0 96.0–100.0
Ba XRF 77 ppm 2.5 98.8–102.6
From geochemical analysis the following conclusions regarding the Co XRF 9 ppm 8.1 93.7–95.4
Cr XRF 17 ppm 2.9 98.4–104.4
depositional environment of C/T black shales at Demerara Rise during
Cu XRF 21 ppm 6.6 94.0–102.2
OAE 2 can be drawn: Mn XRF 78 ppm 2.2 97.5–103.1
Mo XRF 14 ppm 6.0 94.1–104.3
• Elevated FeP/FeT, FeT/Al, and Co/Al values within the OAE 2 confirm Ni XRF 14 ppm 3.3 99.1–103.8
euxinic conditions but, at the same time, require a zone where Rb XRF 26 ppm 3.9 92.7–101.8
Sr XRF 120 ppm 9.7 101.0–102.2
reducing but non-sulfidic conditions enable Fe and Co mobilization
U XRF 5 ppm 10.9 109.0–110.2
in oxygen depleted nearshore sediments. V XRF 56 ppm 4.2 99.8–100.9
• The existence of an expanded oxygen-minimum-zone (OMZ) is Y XRF 8 ppm 5.4 98.5–109.5
demonstrated by extremely low Mn/Al ratios. Elevated Mn/Al ratios Zn XRF 33 ppm 3.9 95.1–102.4
are only seen in distinct horizons. Overgrowth of Mn-carbonates on Zr XRF 30 ppm 3.6 96.1–103.7

pre-existing carbonate tests seems to be possible under euxinic For stable isotope measurements, see section on Materials and methods.
conditions and elevated pore water alkalinities. Notes: Procedures and accuracy of all methods were checked with in-house reference
materials (IR-Analyzer: PS-S (n=286); coulometry: Loess (n=387); XRF: DR-BS (n=33), TW-
• Mo/Al, V/Al and Zn/Al ratios are generally high throughout the
TUC (n=14) and PS-S, (n=11)). Quantification limit is defined as 6 times the maximum
investigated sequence, but a significant depletion is notable exactly standard deviation (1σ). Precision is defined as 100% times the best estimate standard
during OAE 2. The decrease in enrichment for these seawater derived deviation (1σ) divided by the mean of the repeats. The maximum value is given. SD=standard
trace metals may reflect the global onset of black shale deposition deviation. Accuracy is defined as 100% times the mean of the repeat analyses divided by the
during OAE 2 and the rapid drawdown of the seawater TM reservoir expected value. For XRF the range of different reference material is given.

by the enlargement of euxinic depositional areas.


• The enrichment in phosphorus and barium points towards high bio- Appendix B. Identification of so-called “ash layers”
productivity like in recent upwelling areas. Due to diagenesis driven by (see text for details)
the reducing conditions a quantitative interpretation is not possible.
• Besides fixation of sulfide as iron sulfide, organic matter acts as an (a) Profiles for Si/Al, Ti/Al and K/Al for Cenomanian/Turonian black
important sulfur trap during early diagenesis. During OAE 2 the amounts shales at Sites 1258 and 1260. Grey: OAE 2; dashed line: “ash layers”.
of both, pyrite and organically bound sulfur increase. A negative shift in (b) Images of sediment intervals (available online from Erbacher et al.,
the stable sulfur isotope curve during OAE 2 indicates changes in the 2004, doi:10.2973/odp.proc.ir.207.2004). Dashed line: “ash layers”.
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 325
326 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

(c) Comparison of average element contents and element/Al ratios of omitted samples Arthur, M.A., Dean, W.E., Pratt, L.M., 1988. Geochemical and climatic effects of increased
(unknown source. secondary overprinted) marine organic carbon burial at the Cenomanian/Turonian boundary. Nature 335 (6192),
714–717.
Omitted samples Average C/T black shales Average shale Berner, R.A., Raiswell, R., 1983. Burial of organic carbon and pyrite sulfur in sediments
(n = 34) Demerara Rise (Wedepohl, 1971, 1991) over Phanerozoic time: a new theory. Geochimica et Cosmochimica Acta 47 (5),
855–862.
ST % 6.03 2.60 0.24
Bernstein, R.E., Byrne, R.H., 2004. Acantharians and marine barite. Marine Chemistry
SP % (n = 11) 1.67 0.79
86 (1–2), 45–50.
δ34SP ‰ (n = 4) −15.4 − 15.8
Bertram, M.A., Cowen, J.P., 1997. Morphological and compositional evidence for biotic
TIC % 2.73 5.56 0.35 precipitation of marine barite. Journal of Marine Research 55 (3), 577–593.
TOC % 8.55 10.39 Bishop, J.K.B., 1988. The barite–opal–organic carbon association in oceanic particulate
Si % 14.69 10.26 27.53 matter. Nature 332 (6162), 341–343.
Ti % 0.36 0.10 0.46 Bjerrum, C.J., Canfield, D.E., 2002. Ocean productivity before about 1.9 Gyr ago limited
Al % 4.60 2.16 8.84 by phosphorus adsorption onto iron oxides. Nature 417 (6885), 159–162.
FeT % 4.36 1.27 4.80 Böning, P., Brumsack, H.-J., Böttcher, M.E., Schnetger, B., Kriete, C., Kallmeyer, J.,
Mg % 1.01 0.51 1.60 Borchers, S.L., 2004. Geochemistry of Peruvian near-surface sediments. Geochimica
Ca % 10.26 19.61 1.57 et Cosmochimica Acta 68 (21), 4429–4451.
Na % 1.25 0.88 1.19 Böttcher, M.E., 1997. The transformation of aragonite to MnxCa(1 − x)CO3 solid-solutions
K % 0.74 0.64 2.99 at 20 °C: an experimental study. Marine Chemistry 57 (1–2), 97–106.
P % 0.08 0.33 0.07 Böttcher, M.E., Rinna, J., Warning, B., Wehausen, R., Howell, M.W., Schnetger, B., Stein, R.,
Brumsack, H.-J., Rullkötter, J., 2003. Geochemistry of sediments from the
As ppm 54 20 10
connection between the western and eastern Mediterranean Sea (Strait of Sicily,
Ba ppm 511 524 580
ODP Site 963). Palaeogeography, Palaeoclimatology, Palaeoecology 190, 165–194.
Co ppm 23 6 19 Böttcher, M.E., Jørgensen, B.B., Kallmeyer, J., Wehausen, R., 2004. S and O isotope
Cr ppm 136 113 90 fractionation in the western Black Sea. Geochimica et Cosmochimica Acta 68, A345.
Cu ppm 152 63 45 Böttcher, M.E., Hetzel, A., Brumsack, H.-J., Schipper, A., 2006. Sulfur–iron–carbon
Mn ppm 332 17 850 geochemistry in sediments of the Demerara Rise. In: Mosher, D.C., Erbacher, J.,
Mo ppm 32 58 1 Malone, M.J. (Eds.), Proceedings of the Ocean Drilling Program. Scientific Results,
Ni ppm 132 131 68 vol. 207. Ocean Drilling Program, College Station, TX, pp. 1–23.
Rb ppm 22 25 140 Boyle, E.A., 1983. Manganese carbonate overgrowths on foraminifera tests. Geochimica
Sr ppm 564 771 300 et Cosmochimica Acta 47 (10), 1815–1819.
U ppm 7 15 4 Bréhéret, J.G., Brumsack, H.-J., 2000. Barite concretions as evidence of pauses in
V ppm 530 907 130 sedimentation in the Marnes Bleues Formation of the Vocontian Basin (SE France).
Y ppm 17 20 41 Sedimentary Geology 130 (3–4), 205–228.
Breit, G.N., Wanty, R.B., 1991. Vanadium accumulation in carbonaceous rocks. A review of
Zn ppm 209 608 95
geochemical controls during deposition and diagenesis. Chemical Geology 91 (1–2),
Zr ppm 35 33 160
83–97.
ST cf % 12.10 4.94 0.25
Broecker, W.S., Peng, T.-H., 1982. Tracers in the Sea. Eldigio Press, Palisades, New York.
TOCcf % 2.27 19.86 Brumsack, H.-J., 1980. Geochemistry of Cretaceous black shales from the Atlantic Ocean
ST/Al 32.18 1.35 0.03 (DSDP Legs 11, 14, 36, and 41). Chemical Geology 31, 1–25.
TOC/Al 89.00 5.40 0.02 Brumsack, H.-J., 1986. The inorganic geochemistry of Cretaceous black shales (DSDP Leg 41)
Si/Al 3.18 5.13 in comparison to modern upwelling sediments from the Gulf of California. In:
Ti/Al 0.08 0.05 0.05 Summerhayes, C.P., Shackleton, N.J. (Eds.), North Atlantic Paleoocenography. Special
FeT/Al 0.96 0.65 0.54 Publication, vol. 21. Geological Society, London, pp. 447–462.
Mg/Al 0.31 0.53 0.18 Brumsack, H.-J.,1988. Rezente, Corg-reiche Sedimente als Schlüssel zum Verständnis fossiler
Ca/Al 12.03 39.54 0.18 Schwarzschiefer. Habilitationsschrift, Georg-August Universität Göttingen, Germany.
Na/Al 0.27 0.47 0.13 Brumsack, H.-J., 1989. Geochemistry of recent TOC-rich sediments from the Gulf of
K/Al 0.17 0.32 0.34 California and the Black Sea. Geologische Rundschau 78 (3), 851–882.
Brumsack, H.-J., 2006. The trace metal content of recent organic carbon-rich sediments:
P/Al 0.02 0.15 0.01
implications for Cretaceous black shale formation. Palaeogeography, Palaeoceano-
As/Al × 10− 4 12.9 11.2 1.1
graphy, Palaeoecology 232 (2–4), 344–361.
Ba/Al × 10− 4 122.6 294.4 65.6
Brumsack, H.-J., Gieskes, J.M., 1983. Interstitial water trace-metal chemistry of laminated
Co/Al × 10− 4 5.1 3.1 2.1 sediments from the Gulf of California, Mexico. Marine Chemistry 14 (1), 89–106.
Cr/Al × 10− 4 30.6 52.3 10.2 Brumsack, H.-J., Heydemann, A., Kühn, V., Rachold, V., Usdowski, E.,1995. Geochemistry and
Cu/Al × 10− 4 32.6 33.8 5.1 mineralogy of Middle Aptian sediments from the Lower Saxony Basin, NW Germany.
Mn/Al × 10− 4 80.4 15.5 96.2 Neues Jahrbuch für Geologie und Paläontologie. Abhandlungen 196 (2), 235–255.
Mo/Al × 10− 4 6.2 31.9 0.1 Brunner, B., Bernasconi, S.M., 2005. A revised isotope fractionation model for
Ni/Al × 10− 4 28.3 67.5 7.7 dissimilatory sulfate reduction in sulfate reducing bacteria. Geochimica et
Rb/Al × 10− 4 4.8 11.1 15.8 Cosmochimica Acta 69 (20), 4759–4771.
Sr/Al × 10− 4 197.9 693.6 33.9 Calvert, S.E., Pedersen, T.F., 1993. Geochemistry of recent oxic and anoxic marine
U/Al × 10− 4 2.4 11.3 0.4 sediments: implications for the geological record. Marine Geology 113 (1–2), 67–88.
V/Al × 10− 4 119.0 498.0 14.7 Canfield, D.E., Raiswell, R., Westrich, J.T., Reaves, C.M., Berner, R.A., 1986. The use of
Y/Al × 10− 4 5.2 11.7 4.6 chromium reduction in the analysis of reduced inorganic sulfur in sediments and
shale. Chemical Geology 54 (1–2), 149–155.
Zn/Al × 10− 4 49.4 284.4 10.7
Church, T.M., Wohlgemuth, K., 1972. Marine barite saturation. Earth and Planetary
Zr/Al × 10− 4 10.4 20.1 18.1
Science Letters 15 (1), 35–44.
Cenomanian/Turonian black shales of Demerara Rise (mean values of Table 3–5) and Cline, J.D., 1969. Sectrophotometric determination of hydrogen sulfide in natural waters.
‘average shale’. Limnology and Oceanography 14 (3), 454–458.
Crusius, J., Calvert, S., Pedersen, T., Sage, D., 1996. Rhenium and molybdenum
enrichments in sediments as indicators of oxic, suboxic, and sulfidic conditions of
deposition. Earth and Planetary Science Letters 145 (1–4), 66–78.
References Dean, W.E., Arthur, M.A., 1989. Iron–sulfur–carbon-relationship in organic sequences, I.
Cretaceous Western Interior Seaway. American Journal of Science 289 (6), 708–743.
Algeo, T.J., Lyons, T.W., 2006. Mo-total organic carbon covariation in modern anoxic Dymond, J., Suess, E., Lyle, M., 1992. Barium in deep-sea sediments: a geochemical proxy
marine environments: implications for analysis of paleoredox and paleohydro- for paleoproductivity. Paleoceanography 7 (2), 163–181.
graphic conditions. Paleoceanography 21, PA1016. Eagle, M., Paytan, A., Arrigo, K.R., van Dijken, G., Murray, R.W., 2003. A comparison
Algeo, T.J., Maynard, J.B., 2004. Trace element behavior and redox facies in core shales of between excess barium and barite as indicators of carbon export. Paleoceanography
the Upper Pennsylvanian Kansas-type cyclothems. Chemical Geology 206 (3–4), 18 (1), 1021.
289–318. Emerson, S.R., Huested, S.S., 1991. Ocean anoxia and the concentrations of molybdenum
Arndt, S., Brumsack, H.-J., Wirtz, K., 2006. Cretaceous black shales as active bioreactors: and vanadium in seawater. Marine Chemistry 34 (3–4), 177–196.
a biogeochemical model for the deep biosphere encountered during ODP Leg 207 Erbacher, J., Mosher, D.C., Malone, M.J., Shipboard Scientific Party, 2004. Proceedings of
(Demerara Rise). Geochimica et Cosmochimica Acta 70 (2), 408–425. the Ocean Drilling Program. Initial Reports, vol. 207. Ocean Drilling Program,
Arthur, M.A., Schlanger, S.O., Jenkyns, H.C., 1987. The Cenomanian/Turonian Oceanic College Station, TX, p. 89.
Anoxic Event, II. Palaeoceanographic controls on organic-matter production and Erbacher, J., Friedrich, O., Wilson, P.A., Birch, H., Mutterlose, J., 2005. Stable organic
preservation. In: Brooks, J., Fleet, A.J. (Eds.), Marine Petroleum Source Rocks. Special carbon isotope stratigraphy across Oceanic Anoxic Event 2 of Demerara Rise,
Publication, 26. Geological Society, London, pp. 401–420. Western Tropical Atlantic. Geochemistry, Geophysics, Geosystems 6, Q06010.
A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328 327

Forster, A., Schouten, S., Moriya, K., Wilson, P., Sinninghe Damsté, J.S., 2007. Tropical Musavu-Moussavou, B., Danelian, T., 2006. The Radiolarian biotic response to Oceanic
warming and intermittent cooling during the Cenomanian/Turonian oceanic anoxic Anoxic Event 2 in the southern part of the Northern proto-Atlantic (Demerara Rise,
event 2: sea surface temperature records from the equatorial Atlantic. Paleoceano- ODP Leg 207). Revue de Micropaléontologie 49 (3), 141–163.
graphy 22, PA1219. Nederbragt, A.J., Thurow, J., Vonhof, H., Brumsack, H.-J., 2004. Modelling oceanic carbon
Fossing, H., Jørgensen, B.B., 1989. Measurement of bacterial sulfate reduction in and phosphorus fluxes: implications for the cause of the late Cenomanian Oceanic
sediments. Evaluation of a single-step chromium reduction method. Biogeochem- Anoxic Event (OAE2). Journal of the Geological Society 161 (4), 721–728.
istry 8 (3), 205–222. Nederbragt, A.J., Thurow, J., Pearce, R., 2007. Sediment composition and cyclicity in the
Friedrich, O., Erbacher, J., 2006. Benthic foraminiferal assemblages from Demerara Rise mid-Cretaceous at Demerara Rise, ODP Leg 207. In: Mosher, D.C., Erbacher, J.,
(ODP Leg 207, western tropical Atlantic): possible evidence for a progressive Malone, M.J. (Eds.), Proceedings of the Ocean Drilling Program. Scientific Results,
opening of the equatorial Atlantic gateway. Cretaceous Research 27 (3), 377–397. vol. 207. Ocean Drilling Program, College Station, TX, pp. 1–31.
Friedrich, O., Erbacher, J., Mutterlose, J., 2006. Paleoenvironmental changes across the Neretin, L.N., Böttcher, M.E., Jørgensen, B.B., Volkov, I.I., Lüschen, H., Hilgenfeldt, K., 2004.
Cenomanian/Turonian Boundary Event (Oceanic Anoxic Event 2) as indicated by Pyritization processes and greigite formation in the advancing sulfidization front in
benthic foraminifera from the Demerara Rise (ODP Leg 207). Revue de the Upper Pleistocene sediments of the Black Sea. Geochimica et Cosmochimica Acta
Micropaléontologie 49 (3), 121–139. 68 (9), 2081–2093.
Gale, A.S., Jenkyns, H.C., Kennedy, W.J., Corfield, R.M., 1993. Chemostratigraphy versus Neubert, N., Nägler, T., Böttcher, M.E., 2008. Sulfidity controls molybdenum isotope
biostratigraphy: data from around the Cenomanian–Turonian boundary. Journal of discrimination into euxinic sediments: evidence from the modern Black Sea.
the Geological Society 150 (1), 29–32. Geology 36 (10), 775–778.
Galimov, E., 2006. Isotope organic geochemistry. In: Bouillon, S., Böttcher, M.E. (Eds.), Stable Nijenhuis, I.A., Brumsack, H.-J., de Lange, G.J., 1998. The trace element budget of the eastern
Isotopes in Biogeosciences. Organic Geochemistry, vol. 37(10), pp. 1200–1262. Mediterranean during Pliocene sapropel formation. In: Robertson, A.H.F., Emeis, K.-C.,
Gauthier, D.L., 1987. Isotopic composition of pyrite: relationship to organic matter type Richter, C., Camerlenghi, A. (Eds.), Proceedings of the Ocean Drilling Program. Scientific
and iron availability in some North American Cretaceous shales. Chemical Geology Results, vol. 160. Ocean Drilling Program, College Station, TX, pp. 199–206.
65, 293–303. Passier, H.F., Bosch, H.-J., Nijenhuis, I.A., Lourens, L.J., Böttcher, M.E., Leenders, A.,
Gendron, A., Silverberg, N., Sundby, B., Lebel, J.,1986. Early diagenesis of cadmium and cobalt Damste, J.S.S., De Lange, G.J., De Leeuw, J.W., 1999a. Sulfidic Mediterranean surface
in sediments of the Laurentian Trough. Geochimica et Cosmochimica Acta 50 (1), waters during Pliocene sapropel formation. Nature 397 (6715), 146–149.
741–747. Passier, H.F., Böttcher, M.E., de Lange, G.J., 1999b. Sulfur enrichment in organic matter of
Gingele, F.X., Kasten, S., 1994. Solid-phase manganese in Southeast Atlantic sediments: eastern Mediterranean sapropels: a study of sulfur isotope partitioning. Aquatic
Implications for the paleoenvironment. Marine Geology 121 (3–4), 317–332. Geochemistry 5 (1), 99–118.
Hardas, P., Mutterlose, J., 2007. Calcareous nannofossil assemblages of Oceanic Anoxic Paytan, A., Kastner, M., Chavez, F.P., 1996. Glacial to interglacial fluctuations in
Event 2 in the equatorial Atlantic: evidence of an eutrophication event. Marine productivity in the equatorial Pacific as indicated by marine barite. Science 274
Micropaleontology 66 (1), 52–69. (5291), 1355–1357.
Hartmann, M., Nielsen, H., 1969. δ34S-Werte in rezenten Meeressedimenten und ihre Paytan, A., Mearon, S., Cobb, K., Kastner, M., 2002. Origin of marine barite deposits: Sr
Deutung am Beispiel einiger Sedimentprofile aus der westlichen Ostsee. Geolo- and S isotope characterization. Geology, 30 (8), 747–750.
gische Rundschau 58 (2), 621–655. Paytan, A., Kastner, M., Campbell, D., Thiemens, M.H., 2004. Seawater sulfur isotope
Hatch, J.R., Leventhal, J.S., 1992. Relationship between inferred redox potential of the fluctuations in the Cretaceous. Science 304 (5677), 1663–1665.
depositional environment and geochemistry of the Upper Pennsylvanian (Mis- Piper, D.Z., 1994. Seawater as the source of minor elements in black shales, phosphorites
sourian) Stark Shale Member of the Dennis Limestone, Wabaunsee County, Kansas, and other sedimentary rocks. Chemical Geology 114 (1–2), 95–114.
USA. In: Meyers, P.A., Pratt, L.M., Nagy, B. (Eds.), Geochemistry of Metalliferous Black Poulton, S.W., Raiswell, R., 2002. The low-temperature geochemical cycle of iron: from
Shales. Chemical Geology, vol. 99(1–3), pp. 65–82. continental fluxes to marine sediment deposition. American Journal of Science 302 (9),
Hebting, Y., Schaeffer, P., Behrens, A., Adam, P., Schmitt, G., Schneckenburger, P., 774–805.
Bernasconi, S.M., Albrecht, P., 2006. Biomarker evidence for a major preservation Prakash Babu, C., Brumsack, H.-J., Schnetger, B., 1999. Distribution of organic carbon in
pathway of sedimentary organic carbon. Science 312 (5780), 1627–1631. surface sediments along the eastern Arabian Sea: a revisit. Marine Geology 162 (1),
Heggie, D., Lewis, T., 1984. Cobalt in pore waters of marine sediments. Nature 311 91–103.
(5985), 453–455. Raiswell, R., Canfield, D.E., 1998. Sources of iron for pyrite formation in marine sediments.
Helly, J.J., Levin, L.A., 2004. Global distribution of naturally occurring marine hypoxia on American Journal of Science 298 (3), 219–245.
continental margins. Deep-Sea Research Part I 51 (9), 1159–1168. Redfield, A.C., 1958. The biological control of chemical factors in the environment.
Helz, G.R., Miller, C.V., Charnock, J.M., Mosselmans, J.F.W., Pattrick, R.A.D., Garner, C.D., American Scientist 46, 205–221.
Vaughan, D.J., 1996. Mechanism of molybdenum removal from the sea and its Schlanger, S.O., Jenkyns, H.C., 1976. Cretaceous oceanic anoxic events: causes and
concentration in black shales: EXAFS evidence. Geochimica et Cosmochimica Acta consequences. Geologie en Mijnbouw 55, 179–184.
60 (19), 3631–3642. Schlanger, S.O., Arthur, M.A., Jenkyns, H.C., Scholle, P.A., 1987. The Cenomanian–Turonian
Hetzel, A., Brumsack, H.-J., Schnetger, B., Böttcher, M.E., 2006. Inorganic geochemical Oceanic Anoxic Event, I. Stratigraphy and distribution of organic carbon-rich beds
characterization of lithologic units recovered during ODP Leg 207 (Demerara Rise). In: and the marine δ13C excursion. In: Brooks, J., Fleet, A.J. (Eds.), Marine Petroleum
Mosher, D.C., Erbacher, J., Malone, M.J. (Eds.), Proceedings of the Ocean Drilling Source Rocks. Geological Society, London. Special Publication, vol. 26, pp. 371–399.
Program, Scientific Results, 207. Ocean Drilling Program, College Station, TX, pp. 1–37. Schmitz, B., 1987. Barium, equatorial high productivity, and the northward wandering of
Ingall, E.D., Bustin, R.M., Van Cappellen, P., 1993. Influence of water column anoxia on the Indian continent. Paleoceanography 2 (1), 63–77.
the burial and preservation of carbon and phosphorus in marine shales. Geochimica Shipboard Scientific Party, 2004. Leg 207 summary. In: Erbacher, J., Mosher, D.C.,
et Cosmochimica Acta 57 (2), 303–316. Malone, M.J., et al. (Eds.), Proceedings of the Ocean Drilling Program. Initial Reports,
Jacobs, L., Emerson, S., Skei, J., 1985. Partitioning and transport of metals across the O2/ vol. 207. Ocean Drilling Program, College Station, TX, pp. 1–89.
H2S interface in a permanently anoxic basin: Framvaren Fjord, Norway. Geochimica Strauss, H., 1999. Geological evolution from isotope proxy signals: sulfur. Chemical
et Cosmochimica Acta 49 (6), 1433–1444. Geology 161 (1–3), 89–101.
Jacobs, L., Emerson, S., Huested, S.S., 1987. Trace metal geochemistry in the Cariaco Thurow, J., Brumsack, H.-J., Rullkötter, J., Littke, R., Meyers, P., 1992. The Cenomanian/
Trench. Deep-Sea Research 34 (5–6), 965–981. Turonian boundary event in the Indian Ocean — a key to understand the global
Kolonic, S., Sinninghe Damsté, J.S., Böttcher, M.E., Kuypers, M., Kuhnt, W., Beckmann, B., picture. In: Duncan, R.A., Rea, D.K., Kidd, R.B., von Rad, U., Weissel, J.K. (Eds.), The
Scheeder, G., Wagner, T., 2002. Geochemical characterization of Cenomanian/ Indian Ocean: A Synthesis of Results from the Ocean Drilling Program. Geophysical
Turonian Black Shales from the Tarfaya Basin (SW Morocco): relationships between Monograph, vol. 70. American Geophysical Union, pp. 253–273.
palaeoenvironmental conditions and early sulphurization of sedimentary organic Torres, M.E., Brumsack, H.-J., Bohrmann, G., Emeis, K.C., 1996. Barite fronts in
matter. Journal of Petroleum Geology 25 (3), 325–350. continental margin sediments: a new look at barium remobilization in the zone
Kuhnt, W., Herbin, J.P., Thurow, J., Wiedmann, J., 1990. Distribution of Cenomanian– of sulfate reduction and formation of heavy barites in diagenetic fronts. Chemical
Turonian organic facies in the Western Mediterranean and along the adjacent Geology 127 (1–3), 125–139.
Atlantic margin. In: Huc, A.Y. (Ed.), Deposition of Organic Facies. AAPG Studies in Tribovillard, N., Riboulleau, A., Lyons, T., Baudin, F., 2004. Enhanced trapping of
Geology, vol. 30, pp. 133–160. molybdenum by sulfurized marine organic matter of marine origin in Mesozoic
Kuypers, M.M.M., Pancost, R.D., Nijenhuis, I.A., Sinninghe Damsté, J.S., 2002. Enhanced limestones and shales. Chemical Geology 213 (4), 385–401.
productivity led to increased organic carbon burial in the euxinic North Atlantic basin Tribovillard, N., Algeo, T.J., Lyons, T., Riboulleau, A., 2006. Trace metals as paleoredox and
during the late Cenomanian oceanic anoxic event. Paleoceanography 17 (4), 1051. paleoproductivity proxies: an update. Chemical Geology 232 (1–2), 12–32.
Lyons, T.W., Berner, R.A., 1992. Carbon–sulfur–iron systematics of the uppermost deep van der Weijden, C.V., 2002. Pitfalls of normalization of marine geochemical data using
water sediments of the Black Sea. Chemical Geology 99 (1–3), 1–27. a common divisor. Marine Geology 184 (3–4), 167–187.
Lyons, T.W., Severmann, S., 2006. A critical look at iron paleoredox proxies: new insights from Wanty, R.B., Goldhaber, M.B., 1992. Thermodynamics and kinetics of reactions involving
modern euxinic marine basins. Geochimica et Cosmochimica Acta 70 (23), 5698–5722. vanadium in natural systems — accumulation of vanadium in sedimentary rocks.
McManus, J., Berelson, W.M., Klinkhammer, G.P., Johnson, K.S., Coale, K.H., Anderson, R. Geochimica et Cosmochimica Acta 56 (4), 1471–1483.
F., Kumar, N., Burdige, D.J., Hammond, D.E., Brumsack, H.-J., McCorkle, D.C., Rushdi, Wedepohl, K.H., 1971. Environmental influences on the chemical composition of shales
A., 1998. Geochemistry of barium in marine sediments: implications for its use as a and clays. In: Ahrens, L.H., Press, F., Runcorn, S.K., Urey, H.C. (Eds.), Physics and
paleoproxy. Geochimica et Cosmochimica Acta 62 (21–22), 3453–3473. Chemistry of the Earth, vol. 8. Pergamon, Oxford, pp. 305–333.
Morford, J.L., Emerson, S.E., 1999. The geochemistry of redox sensitive trace metals in Wedepohl, K.H., 1991. The composition of the upper earth's crust and the natural cycles of
sediments. Geochimica et Cosmochimica Acta 63 (11–12), 1735–1750. selected metals. Metals in natural raw materials. In: Merian, E. (Ed.), Metals and Their
Mort, H.P., Adatte, T., Föllmi, K.B., Keller, G., Steinmann, P., Matera, V., Berner, Z., Stüben, Compounds in the Environment. Natural Resources. VCH, Weinheim, pp. 3–17.
D., 2007. Phosphorus and the roles of productivity and nutrient recycling during Werne, J.P., Hollander, D.J., Lyons, T.W., Sinninghe Damsté, J.S., 2004. Organic sulfur
oceanic anoxic event 2. Geology 35 (6), 483–486. biogeochemistry: recent advances and future research directions. In: Amend, J.,
328 A. Hetzel et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 302–328

Edwards, K., Lyons, T. (Eds.), Sulfur Biogeochemistry: Past and Present. Special Wortmann, U.G., Bernasconi, S.M., Böttcher, M.E., 2001. Hypersulfidic deep biosphere
Paper, vol. 379. Geological Society of America, pp. 135–150. indicates extreme sulfur isotope fractionation during single step microbial sulfate
Wijsman, J.W.M., Middelburg, J.J., Herrmann, P.M., Böttcher, M.E., Heip, C.H.R., 2001. reduction. Geology 29 (7), 647–650.
Sulfur and iron speciation in surface sediments along the northwestern margin of Zhabina, N.N., Volkov, I.I., 1978. A method of determination of various sulfur compounds
the Black Sea. Marine Chemistry 74 (4), 261–278. in sea sediments and rocks. In: Krumbein, W.E. (Ed.), Environmental Biogeochem-
Wortmann, U.G., Chernyavsky, B., 2007. Effect of evaporite deposition on Early istry and Geomicrobiology (vol. 3): Methods, Metals and Assessment. Ann Arbor
Cretaceous carbon and sulphur cycling. Nature 446 (7136), 654–656. Science Publications, pp. 735–745.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 329

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Abstract of “Climate-Ocean coupling off North-West Africa during the Lower Albian:
The Oceanic Anoxic Event 1b”☆
Peter Hofmann a,⁎, Isabel Stüsser a, Thomas Wagner b, Stefan Schouten c, Jaap S. Sinninghe Damsté c
a
Institute for Geology and Mineralogy, University of Cologne, Zülpicher Str. 49a, 50674 Köln, Germany
b
School of Civil Engineering and Geosciences, Newcastle University, Newcastle upon Tyne, NE1 7RU, United Kingdom
c
Department of Marine Organic Biogeochemistry, NIOZ Royal Netherlands Institute for Sea Research, Post Office Box 59, 1790 Den Burg, The Netherlands

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

Keywords: High resolution records of organic and inorganic geochemical proxies document environmental changes
Oceanic Anoxic Event 1b during Oceanic Anoxic Event 1b for the Albian upwelling system at DSDP Site 545 Mazagan Plateau off NW
Black shale Africa. Sea surface temperature estimates based on TEX86 analyses show an abrupt rise by ~ 3 °C concurrent
Sea surface temperatures
with a more than 2-fold increase in accumulation rates for organic carbon and siliciclastic sediment
Upwelling system
components. Geochemical grain-size indicators (Si/Al and Zr/Al) indicate the influx of overall finer grained
sediment during the event. Our data suggest that the warming phase associated with OAE 1b resulted in an
attenuation of the NE trade wind system and a weakening of local upwelling conditions. Higher amounts of
precipitation over NW Africa and an associated higher continental run off may have supplied nutrients to the
eastern North Atlantic and fostered the production of organic matter and black shale deposition.
© 2008 Elsevier B.V. All rights reserved.

☆ The full version of this paper has been published in: Palaeogeography, Palaeoclimatology, Palaeoecology 262, 2008, 157–165.
DOI of original article: 10.1016/j.palaeo.2008.02.014.
⁎ Corresponding author. Fax: +49 221 4705149.
E-mail address: peter.hofmann@uni-koeln.de (P. Hofmann).

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.12.011
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Temperature controlled deposition of early Cretaceous (Barremian–early Aptian)


black shales in an epicontinental sea
Jörg Mutterlose a,⁎, Sebastian Pauly a, Thomas Steuber b
a
Institut für Geologie, Mineralogie und Geophysik, Ruhr-Universität Bochum, Universitätsstr. 150, 44801 Bochum, Germany
b
The Petroleum Institute, PO Box 2533, Abu Dhabi, UAE

A R T I C L E I N F O A B S T R A C T

Article history: A continuous, 103m thick mudstone sequence of Barremian–early Aptian age, recently exposed in northern
Received 16 November 2007 Germany, has been logged and analyzed with respect to its macrofaunal content and its geochemistry (δ13C,
Received in revised form 6 April 2008 δ18O). Based on common belemnite and rarer ammonite findings a detailed biozonation has been established
Accepted 17 April 2008
covering the complete Barremian–early Aptian interval, including the mid early Aptian “Fischschiefer”, the
deposition of which occurred during Oceanic Anoxic Event 1a. Quite common throughout the sequence are
Keywords:
Laminated sediments
finely laminated black shales (Blätterton) with high TOC contents of up to 7%. These laminated black shales
Anoxic conditions with CaCO3 contents of up to 60% are interbedded with dark and carbonate-poor clays on the scale of
Barremian/Aptian decimeters to several meters. Oxygen depletion of bottom waters is indicated for the laminated intervals
Belemnites since these are barren of any benthic fossils.
Stable isotopes A salinity driven stratification of the water column during more humid periods was caused by a more intense
Cooling run-off from the nearby hinterland. This and increased surface water productivity are seen as a possible cause
for the lamination with observations being mainly based on micropalaeontological evidence (dinoflagellates,
calcispheres, calcareous nannofossils). Stable isotope studies (δ13C, δ18O) performed on belemnite guards
show δ13C values of 0‰ to 2‰ for the Barremian. A distinctive positive excursion of ~ 4‰ directly above the
early Aptian Fischschiefer correlates well with the global positive δ13C excursion of mid early Aptian age
following the Oceanic Anoxic Event 1a. This first record of the δ13C positive excursion from the Boreal Realm
therefore confirms the global nature of the signal. A shift in δ18O values of about 2‰ in the late Barremian–
early Aptian is interpreted as a temperature decrease by ~ 8°C. This cooling goes along with a shift from the
thick (up to 6.5m) early Barremian Hauptblätterton to thin (10cm to 20cm) laminites (Blätterton) in the late
Barremian. During the deposition of the Hauptblätterton very warm conditions caused a long lasting (N
500kyrs) stable water stratification. The cooler temperatures of the late Barremian reduced the stability of
the water stratification resulting in rather short termed alternations of thin laminites and dark clays. These
observations suggest a temperature controlled peak of anoxic conditions in the early Barremian
Hauptblätterton and subsequently short termed changes from anoxic to suboxic conditions in the late
Barremian and early Aptian.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction closed the former sea way to the south via Poland. The restricted
palaeogeographic setting throughout the Barremian and the earliest
Throughout the Barremian and early Aptian the region presently Aptian resulted in the deposition of the finely laminated black shales.
occupied by the North Sea, the Netherlands, northern Germany, and Oxygen depletion of bottom waters is indicated for the laminated
Poland formed the southern extension of a shallow marginal sea that intervals since these are barren of benthic fossils. They yield, however,
was connected to the Boreal–Arctic Sea in the north by an open sea well preserved nektonic and planktonic organisms, as well as
way (Fig. 1). This semi-restricted basin was characterized by the terrestrial plant material (Mutterlose and Böckel, 1998; Mutterlose
deposition of more than 200m of clay- and mudstones with distinctive and Bornemann, 2000).
laminated clay horizons (“Hauptblätterton”, “Blätterton”, “Fischschie- The origin and genesis of these Barremian black shales have been
fer”) occurring throughout the succession. These sediments rich in C- widely discussed in recent years (Mutterlose and Böckel, 1998). A
org, reflect basin-wide anoxic to suboxic conditions, caused by a major stratification of the water column due to periodically increased run-off
regression in latest Hauterivian to earliest Barremian times, which has been suggested as a possible cause with observations being mainly
based on micropalaeontological evidence (dinoflagellates, calci-
⁎ Corresponding author. spheres, calcareous nannofossils). The most prominent of these
E-mail address: joerg.mutterlose@rub.de (J. Mutterlose). laminated horizons, the 5m to 6m thick Hauptblätterton of late

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.04.026
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 331

Fig. 1. Paleogeographic map of the Barremian of NW Europe showing the sections and localities mentioned in the text. DB = Danish Basin, sNSB = southern North Sea Basin, cNB =
central Netherland Basin, LSB = Lower Saxony Basin, PB = Polish Basin. 1 = A39 section; 2 = cores Hoheneggelsen KB 40, KB 50; 3 = Gott and Moorberg sections.

early Barremian age, has recently been the subject of a detailed continuous subsidence, and great thicknesses of shales and mud-
geochemical study (δ13C, δ18O, TEX86, trace elements) interpreting this stones accumulated. Shallow water sediments, including sands and
marker horizon as a warm water event (Mutterlose, personal iron-rich mudstones, were deposited along the margins of the basin in
communication). There are, however, no stable isotope data available the west, south and east. The evolution of the basin has been
for the complete early Barremian–early Aptian interval, which might described in detail (Schott et al., 1969; Michael, 1974, 1979; Kemper,
shed light on the long-term climatic trend, which in turn may be used 1979; Mutterlose, 1992; Mutterlose and Bornemann, 2000).
to decipher the origin of the laminites. After a final transgressive peak in the latest Hauterivian [Simbirs-
A temporary road cutting in 2005 offered the opportunity to log kites (Craspedodiscus) discofalcatus ammonite Zone] significant
and sample a 103m thick fossiliferous succession covering the palaeogeographic and palaeoceanographic changes occurred at the
complete lower Barremian–lower Aptian interval. Common belemnite Hauterivian/Barremian boundary, and the overall character of the
and rarer ammonite findings allowed for a detailed biozonation which Barremian succession in NW Europe is regressive (Rawson and Riley,
formed the base for further detailed isotope studies. The current paper 1982; Ruffell, 1991). A regression in the earliest Barremian led to
aims at reconstructing the palaeotemperature record for the Boreal brackish–lacustrine conditions in central and southern Poland (Marek
lower Barremian–lower Aptian by supplying stable isotope data (δ13C, and Racynska, 1979), isolating the North Sea and adjoining areas from
δ18O) from belemnite guards. Further on we want to use these findings a marine sea way to the south.
to better understand the genesis of the laminated clays and to describe A thick laminated marker horizon, the Hauptblätterton of late early
the environmental changes that occurred in the Barremian–early Barremian age, marks the establishment of anoxic conditions
Aptian seas. throughout the Lower Saxony Basin, the remainder of the overlying
Barremian sequence consisting of only two main lithologies. Dark
2. Geological setting clays and laminated marls form a clay–Blätterton alternation (Fig. 2).
The timing of the onset and the thickness variation of the Barremian
2.1. Sediments laminites allows for a differentiation of a basin centre and a basin
margin. In the basin centre earliest laminites occur already near the
Marine sediments of early Barremian–early Aptian age have a Hauterivian–Barremian boundary interval, at the basin margin they
widespread distribution in northern Germany, the North Sea and to a first appear in the late early Barremian (Hauptblätterton). While thin,
certain extent in northeast England, where they have been studied for i.e. 10cm to 20cm thick laminites (Blättertone) are quite common
over a hundred years. During this period NW Europe was composed of throughout the late Barremian of the basin centre, they are much rarer
a number of basins (central Netherland Basin, Lower Saxony Basin, at the basin margin. These patterns suggest that anoxic conditions
Danish–Polish Basin, southern North Sea Basin), forming the southern prevailed in the basin centre from latest Hauterivian time onward,
extension of the Boreal–Arctic Sea further to the north (Fig. 1). expanding onto the margins only in the late early Barremian (Fig. 3).
Seaways towards the Tethys in the south did not exist during this Detailed stratigraphical and geochemical data of the clay–
period. The Lower Saxony Basin formed the southernmost of these Blätterton alternation have been compiled by Benesch et al. (1998),
basins, having a length of 280km and a width of 80km. In Barremian Bischoff and Mutterlose (1998), Hild and Brumsack (1998) and Littke
and Aptian times the central parts of this basin were characterized by et al. (1998). The late Barremian dark clays contain an average of 2%
332 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

Fig. 2. Bio- and lithostratigraphy of the Barremian–Aptian in northern Germany.

Corg and they are poor in calcium carbonate which varies from 0% to Hedbergella Marls can be subdivided into a lower (Neohibolites ewaldi
10% and decreases toward the top of the Barremian. The latest Marl), a middle (Neohibolites clava Marl) and an upper (Neohibolites
Barremian depressa Clay (Fig. 2) has calcium carbonate values of 0% to inflexus Marl) unit.
1%, being rich in pyrite and siderite. The laminites throughout the
Barremian are enriched in organic matter (2% to 7% TOC), their calcium 2.2. Palaeontology
carbonate content ranges from 10% to 45%. The organic particles are
mainly composed of liptinites that formed from spores, pollen and While benthic organisms (foraminifera, macrofossils) are in
algae. The liptinites increase toward the top, with the Fischschiefer general rare throughout the Barremian–lower Aptian mudstone
being the most liptinite-enriched end member (90% to 95%) of the sequence, phytoplankton (calcareous nannofossils, calcispheres, dino-
anoxic sequence. Terrigenous organic matter (vitrinite and inertinite) flagellates) and nektonic faunas (ammonites, belemnites, fish) are
on the other hand is more common in the clay–Blätterton alternation. quite common. The palaeogeographic and palaeoceanographic con-
The restricted palaeoceanographic conditions of the Barremian which figuration (Fig. 1) favoured the evolution of endemic species from
allowed the accumulation of these organic-rich sediments prevailed Tethyan ancestors. Among both nannofloras and faunas, endemic
into earliest Aptian time, as dark clays (Deshayesites bodei Clay) are species evolved simultaneously and became quite abundant. These are
overlain by the Fischschiefer rich both in organic matter and calcium associated with Boreal elements, whereas Tethyan forms are extre-
carbonate (Fig. 2). Above the Fischschiefer, pale marls indicate a change mely rare or absent.
in the depositional environment to hemipelagic conditions. Named after Calcareous nannofossils are common in the lower Barremian, their
the common planktonic foraminifera Hedbergella, these Hedbergella diversity and abundance decrease in the upper Barremian. The dark O.
Marls correspond to the Gargasian Stage of the French subdivision. depressa Clay of latest Barremian age is virtually barren, reflecting the
Based on findings of different species of the belemnite Neohibolites, the overall low CaCO3 content of these beds. Rich and diverse assemblages
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 333

Fig. 3. Lithostratigraphic correlation of laminites of the basin margin and basin centre.

of calcareous nannofossils, consisting of up to 58 species, have been group is absent in most samples of the non-laminated facies. The data
listed from the Blätterton and the Fischschiefer (Mutterlose and gained from palynomorphs thus very clearly indicate a strong
Harding, 1987; Habermann and Mutterlose, 1998). A distinctive terrigenous input, reduced salinity of surface waters, and a weaker
sequence of calcareous nannofossil species on the scale of 10μm, marine signal for the deposition of the laminites of the late Barremian.
described from the late early Barremian laminated Munk Marl Bed In the marginal parts of the basin these phases of increased terrestrial
(= Hauptblätterton) of the Danish Sector of the North Sea by Thomsen input are reflected by silt components. The non-laminated clays and
(1989), implies that climatic/oceanographic changes in the Barremian marls of the early Aptian in contrast record a stronger marine signal.
were effective enough to cause seasonal changes. The calcispheres of the clay–Blätterton alternation are dominated
The non-laminated black clays and marls of the upper Barremian by obliquipithonelloid cysts, with coarse-crystalline forms being more
and lower Aptian are poor in amorphous kerogen and relatively common than fine crystalline ones. These associations indicate cold
impoverished in terrestrial palynomorphs (spores, pollen). On the water conditions in a pelagic regime (Keupp and Mutterlose, 1994).
other hand dinoflagellates show both high diversity and high Barremian benthic foraminifera show a gradual decrease in
abundance within these non-laminated sediments, reflecting a strong diversity and abundance over the succession, with agglutinated
marine signal. An average of 72 species of dinoflagellates has been foraminifera dominating in sediments of latest Barremian and early
encountered (Below and Kirsch, 1997). The laminated sediments are Aptian age; tiny very thin-shelled calcareous foraminifera are rare. The
enriched in amorphous kerogen and in terrestrial palynomorphs. Both Hauptblätterton and Blätterton horizons yield extremely impover-
Blätterton horizons and the Fischschiefer show strong influxes of ished microfaunas; only a few species of ostracods and benthic
terrestrial matter in comparison to the black clays. The abundance and foraminifera are known from these beds (Michael, 1967). Rarely there
diversity of dinoflagellates decrease in these sediments. Averages of are more than 10 species in a sample.
36 species were found in the Blätterton and 32 species in the The cephalopod faunas of the Barremian of NW Europe are
Fischschiefer. Within palynomorphs the Prasinophyta (Pterospermella, dominated by the belemnite family Oxyteuthididae (suborder
Leiosphaera) show quite distinctive fluctuations. This group, which is Belemnitina), a group which is restricted to the Boreal Realm. The
believed to indicate a reduced salinity of surface waters (Below and Oxyteuthididae are extremely common and the most useful index
Kirsch, 1997) is enriched in the laminites. Prasinophyta make up an fossils in the Barremian of NW Europe (Mutterlose, 1990). The species
average of 24% of total organic matter in the Blätterton, whereas this succession of the Oxyteuthididae both in NW Germany and NE
334 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

England clearly shows an evolutionary lineage: Praeoxyteuthis 3. Material


jasikofiana (Hauterivian–Barremian boundary interval)–Praeoxy-
teuthis pugio (earliest Barremian)–Aulacoteuthis spp. (late early Material (bulk rock samples, belemnite guards) has been studied
Barremian)–Oxyteuthis brunsvicensis (early late Barremian)–Oxy- from a road cutting of the motorway “A39” located three km east of
teuthis germanica (mid late Barremian)–Oxyteuthis depressa (latest Braunschweig (northern Germany; coordinates: 52°14.819 N, 10°36.657
Barremian). O. depressa, the last species in the Barremian is the most E to 52°14.845 N,10°36.322 E; Fig.1). Today, the road cutting is no longer
common macrofossil in the dark black clays of the basin facies (Fig. 4). accessible, but when the samples were collected in 2005, a 103m thick
This Praeoxyteuthis–Aulacoteuthis–Oxyteuthis assemblage is endemic sequence of Barremian–early Aptian sediments was exposed. The age
to the southern North Sea (northeast England, northern Germany), it assignment and biozonation of this succession employed here is based
differs from the assemblages of the Russian Platform and is unknown on belemnites and ammonites (see Section 5.1).
from the Tethys. The lower Barremian is characterized by 12.5m of mudstones,
A diverse and excellent preserved fish fauna has been described which include 5.7m of finely laminated CaCO3 rich black shales at its
from the Barremian laminites (Brahms, 1913) supporting the absence top. These finely laminated black shales yield a rich belemnite fauna
of benthic organisms due to oxygen depletion of the bottom waters. (Aulacoteuthis spp.), and are here assigned to the Hauptblätterton of
This diverse fish fauna is dominated by teleostei (13 species), ganoid the Aulacoteuthis spp. belemnite zone (Fig. 5). The thickness of the
fishes are less common (6 species). The Fischschiefer owes its name to upper Barremian (Fig. 5) cannot be fixed exactly, due to a 5.5m thick
the well preserved fossil fishes which have been described from interval barren of index fossils near the Barremian/Aptian boundary. It
Helgoland (Taverne and Ross, 1973). Based on taphonomic observa- varies from 72.5m (last occurrence of the late Barremian belemnite
tions these fishes were deposited in a lagoonal setting. Oxyteuthis depressa in bed 165) to 78m (first occurrence of the early
In conclusion the palaeontological data suggest anoxic (Blätter- Aptian ammonite Deshayesites deshayesi in bed 170). The upper
ton)–suboxic conditions (dark clay) of bottom waters in a semi- Barremian consists of dark clays and laminated mudstones containing
restricted epicontinental sea, characterized by often endemic marine an abundant belemnite fauna assigned to the Oxyteuthis brunsvicensis,
floras and faunas. Variations of run-off caused changes of surface Oxyteuthis germanica and O. depressa zones. The uppermost 12.5m of
water salinity with the laminites indicating a stronger input of the sequence which include dark clays, laminated mudstones, three
terrestrial matter. This implies that the anoxic conditions which tuff horizons and pale marls is of early Aptian age, based on findings of
caused the deposition of the laminites went along with a periodically the belemnite Neohibolites ewaldi and the ammonite D. deshayesi.
stable halocline and increased surface water productivity of the A total of 119 bulk-rock samples, collected bed-by-bed, has been
primary producers, which in turn resulted from seasonally increased studied from the 103m thick sequence for CaCO3, 83 samples for TOC
run-off from the nearby hinterland. (Appendix A). A total of 89 belemnite guards has been collected bed-

Fig. 4. Belemnite zonation of the Boreal Barremian.


J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345
Fig. 5. Litho- and biostratigraphy of the A39 section showing the TOC- and CaCO3 values. P. p. = Praeoxyteuthis pugio, A. = Aulacoteuthis, O. b. = Oxyteuthis brunsvicensis, O. g. = Oxyteuthis germanica, O. d. = Oxyteuthis depressa, N. e. = Neohibolites ewaldi, D.
d. = Deshayesites deshayesi.

335
336 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

Fig. 5 (continued ).
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 337

teuthis stock form an evolutionary lineage which passes from one another
with minor overlap of the individual species. The high abundance of these
taxa throughout the Boreal Barremian allows for a biometric–statistical
analysis of this group, resulting in more objective species criteria rather
than pure morphological descriptions. Praeoxyteuthis includes elongate
slender taxa without groove, Aulacoteuthis is defined by its slender form
and a ventral groove. Oxyteuthis includes rather stout ungrooved forms
with dorsoventral flattening of the guards increasing from O. brunsvicensis
via O. germanica to O. depressa (Fig. 4).

4.2. Bulk rock geochemistry

The CaCO3 content was measured by using the “Müller-Bombe”


(Müller and Gastner, 1971; accuracy ± 3wt.%). The bulk rock samples
(2.0g), which were dried and homogenized, were subsequently
dissolved with HCl (15%) in a glass bottle equipped with a manometer.
The CaCO3 content was calculated by the relative gas pressure increase
inside the glass bottle. TOC data were obtained by combustion of dried
bulk rock samples at ~ 1250°C. The evolving CO2, which corresponds
to the total carbon (TC) was measured coulometrically. The amount of
TOC was then calculated by using the CaCO3 and CO2 data by the
equation: TOC = (TC) − (CaCO3 ⁎ 0.12).

4.3. Belemnite geochemistry

A total of 68 belemnite guards was studied with respect to their trace


elements and stable isotope values (δ13C, δ18O). After boiling the guards
in water all specimens were split dorsoventrally along the “Klähnsche”
interface (see Ernst, 1964) by using a small chisel. Material (2mg) was
then removed from the split surfaces by using a microdrill (Proxxon
hand-held driller, 0.4mm tungsten drillbit). Since the δ13C and δ18O
values of Barremian belemnites do not show a significant ontogenetic
variation for oxyteuthid belemnites, one sample was taken from each
rostrum. The sampling took place halfway between the apical line and
the outer surface. Prior to stable isotope studies, each sample was
analyzed for trace element concentrations (Mg, Mn, Fe, Sr) using an ICP-
AES (SpectroCirosCCD). Owing to errors in weighing of samples, precision
of these results was ± 8% of the reported concentrations. Analytical
detection limits for Mn and Fe were 8ppm and 80ppm, respectively.
Stable isotope values (13C/12C, 18O/16O) were measured using a Finnigan
MAT 251 mass spectrometer combined with a Carbo Kiel device. All
reported δ-values refer to the PDB standard if no other standard is
indicated. Reproducibility of isotopic results (± 0.08‰ δ13C and ± 0.1‰
δ18O) was monitored by continuous analyses of standard materials.
Palaeotemperatures were calculated using the δ18O notation following
the equation of Epstein et al. (1953) andCraig and Gordon (1965), as
modified by Anderson and Arthur (1983):

T ½ C ¼ 16:0−4:14*ðδCaCO3 −δH 2 O½SMOW Þ


þ 0:13*ðδCaCO3 −δH 2 O½SMOW Þ2 :


Fig. 5 (continued ).

In this equation δH2O is the oxygen isotopic composition of


seawater from which the calcite was precipitated, referring to the
by-bed allowing for a detailed biostratigraphic zonation. Out of these, SMOW standard. Since the seawater δ18O value is related to ice
68 guards have been analyzed for their trace elements and volumes and latitudinal gradients, different values have been used.
subsequently for their stable isotope values (δ13C, δ18O). Today's world would ask for a δ18O value of 0‰ SMOW, an ice free
world which has been suggested for the Cretaceous requires a δ18O
4. Methods value of − 1‰ SMOW (Shackleton and Kennett, 1975); both values
have been used in our palaeotemperature calculations (Appendix B).
4.1. Belemnite taxonomy Because of the difficulties in quantifying the impact of potential ice
volumes and latitudinally controlled evaporation in particular in
The taxonomic concept used for the belemnites is that of Mutterlose relative small marginal seas, more recent studies (e.g. McArthur et al.,
(1983), and Mutterlose and Baraboshkin (2003). The species assignments 2007) only use δ18O data without referring to absolute temperatures.
follow morphological aspects with individual species based on biometric We follow this approach, but list the calculated temperatures in
criteria. The various species of the Praeoxyteuthis–Aulacoteuthis–Oxy- parenthesis for both seawater δ18O values (0‰; − 1‰).
338 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

5. Results swimmer living in subsurface waters of perhaps 50m to 150m (for


details see Mutterlose. personal communication).
5.1. Palaeobiology and biostratigraphy Ammonites have been observed in the A39 section from bed 170
about 0.5m below the first Fischschiefer horizon. A total of 31
The most applicable biostratigraphic scheme of the Boreal ammonites has been attributed to the genera Prodeshayesites, Deshaye-
Barremian is based on belemnites, which are quite common and for sites, Aconeceras, Ancyloceras, Sanmartinoceras confirming an early
which a modern taxonomy exists. Out of the 89 guards collected from Aptian age. For a more detailed discussion of the early Aptian ammonite
the A39 section 80 specimens were assigned to the Praeoxyteuthis– faunas see Casey (1964), Kemper (1971, 1995) and Raisossadat (2004).
Aulacoteuthis–Oxyteuthis stock, four to Hibolithes minutus, endemic to
the Boreal Barremian. Five specimens were attributed to Neohibolites 5.2. Bulk rock geochemistry
ewaldi typical for the early Aptian. A detailed list of the species is given
in Appendix B. All zonal markers characteristic for the Barremian were The CaCO3 values vary from 0% to 5% in the dark clays and from 8%
observed implying a complete, undisturbed sequence. Belemnite to 30% in the laminated black shales (Appendix A). These results are
guards were found both in the dark clays and in the laminites, species consistent with previous data, which stated a maximum of 45% CaCO3
of Aulacoteuthis occurring abundantly in the Hauptblätterton (Fig. 5, for the laminated sediments (Mutterlose and Böckel, 1998). Carbonate
Appendix B). These findings of belemnites in sediments of anoxic is mainly contributed by calcareous nannofossils and calcispheres,
origin, which are barren of any benthic organisms, exclude a while both benthic and planktonic foraminifera are absent. The
nektobenthic life style for Aulacoteuthis like that of recent Sepia. It benthic foraminifera are very rare or absent due to the anoxic bottom
rather suggests a nektonic life like recent teuthids. This view is water conditions (Michael, 1967), while planktonic foraminifera did
supported by findings of arm hooks, which indicate a life as an active not yet occur in Barremian times in the Boreal Realm. The basal part of
predator, and TEX86 palaeothermometry data (Mutterlose, personal the overlying Hedbergella Marl is marked by a remarkable change to
communication). A constant offset of warmer TEX86 data and diverse and abundant assemblages of benthic and planktonic
relatively cooler δ18O belemnite signal by 4°C to 5°C during the foraminifera (Rückheim and Mutterlose, 2002).
Hauptblätterton has been observed, which has been interpreted to The TOC values vary from 0.1% to 7%, lowest values were measured
reflect temperature signals from two different depth habitats. in the tuff horizon 179. The TOC content of the Hauptblätterton ranges
According to these interpretations Aulacoteuthis was a permanent from 0.6% to 4.3%, that of the Blättertone from 1.4% to 6.2%.

Fig. 6. Trace elements of belemnites from the A39 section.


J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 339

5.3. Trace elements of belemnites 5.4. Stable isotope data of belemnites

Concentrations of selected trace elements (Mg, Mn, Fe, Sr) Out of the 68 belemnites (Fig. 6, Appendix B) analyzed from the
determined from the belemnites studied are shown in the cross- early Barremian–early Aptian sequence, 64 specimens have been
plots of Fig. 6. Mg concentrations vary from 913ppm to 4898ppm considered for the following interpretation. The four specimens that
(Fig. 6, Appendix B), Mn concentrations from b 8ppm to 135ppm. Most show high Mn and Fe concentrations (Fig. 6) are listed in Appendix B
of the data points range from 1000ppm to 3000ppm for Mg (low-Mg- (marked with an asterisk) but are not considered in Fig. 7.
calcite) and from 1ppm to 60ppm for Mn. For Mn, a threshold of The δ13C values of the belemnites delineate a threefold subdivision
100ppm is considered to be critical for diagenesis (Steuber, 1999). Sr of the section studied. In the lower 30m of the succession, rapid
concentrations vary from 1064ppm to 2238ppm, indicating a pristine changes from 0‰ to 2‰ are recorded within one horizon. Three
signal without diagenetic overprint. The critical threshold indicative belemnite guards from a 10cm thick Blätterton horizon (bed 124) gave
for the latter case is Sr concentrations below 1000ppm (Steuber, a range from − 0.26‰ to + 1.62‰. Belemnites from laminites and clays
2001). A trend of decreasing Sr concentrations and increasing Mn in most cases gave similar results. The middle part from 30m to ~ 80m
concentrations would show a diagenetic alteration. Apart from four can be described as a plateau with most values ranging from + 1‰ to +
belemnites (99/3, 109/2, 130/1, 140 base) all trace elements data 2‰. A significant increase to 4‰ above the early Aptian Fischschiefer
therefore indicate a pristine preservation of the belemnite skeletal (~ 95m) marks the third interval.
calcite and suggest that stable isotope values were not altered by a The δ18O temperature estimates, assuming a seawater value of 0‰
significant diagenetic overprint. SMOW (today) and − 1‰ SMOW (ice free Cretaceous), respectively

Fig. 7. δ13C and δ18O data from belemnites of the A39 section.
340 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

increase from sample 99/2 (17°C; 13°C) below the Hauptblätterton to The positive δ13C excursion directly above the Fischschiefer (94.6m
sample 100/4 (25°C; 20°C) in the basal part of the laminated black to 96.9m) merits attention. Though the Fischschiefer clearly is a
shales with constant values around (25°C; 20°C) in the upper part of regional phenomenon (Mutterlose and Böckel, 1998), it corresponds
the Hauptblätterton. The uppermost part of the Hauptblätterton has stratigraphically to the Oceanic Anoxic Event 1a (= OAE 1a), which
recorded an increase of the δ18O values to − 1.8‰ (sample 100/14; correlates to the Italian “Selli” black shale event. The Selli level
24°C; 19°C). It is in the directly overlying black clays that the δ18O observed in the Cismon section (Italy) and elsewhere is coeval with the
values show an increase to − 0.8‰ (sample 103/1; 20°C; 15°C). The Fischschiefer, both successions showing a similar succession of
remainder of the O. brunsvicensis and O. germanica clay–Blätterton calcareous nannofossil events. In Italy, the Selli level is sandwiched
alternation (12.5m to 30m above base) gave less negative δ18O values by the first occurrences (= FOs) of Chiastozygus litterarius/Rucinolithus
of − 1.25‰ (115 mid; 21°C; 17°C) to 0‰ (112 base; 16°C; 12°C). These irregularis and the FOs of Rhagodiscus angustus/Eprolithus floralis (Erba,
show considerable fluctuations; within 20cm the δ18O values 1994). In northern Germany R. angustus has its FO within the
decreases by nearly 0.8‰ from − 0.94‰ to − 0.17‰. The upper part Fischschiefer while the FOs of E. apertior and E. floralis occur just
of the succession (30m to 103m above base), attributed to the upper above this black shale (Bischoff and Mutterlose, 1998). The positive
part of the O. germanica zone, the O. depressa zone and the N. ewaldi δ13C excursion to 4‰ directly above the Fischschiefer which is based on
zone, reveals a trend to values around ~ 0‰ (152/1; 16°C; 12°C). While three belemnite guards corresponds to segment C6 and the lower part
the δ18O values still vary from ~ 0‰ to − 1.3‰ (2m above 149; 21°C; of C7 (sensu Menegatti et al., 1998) observed in the Cismon section,
17°C) from 30m to 48m, the trend of the interval 48m to 103m is where the δ13Ccarb data (bulk rock) increase up to 4‰ in limestones
rather stable with only minor variations. overlying the Selli level, below the FO of E. floralis (Weissert et al., 1998;
Stable isotope patterns are often reflecting diagenetic alteration. In Hochuli et al., 1999). This event records a global perturbation of the
our case the δ13C and δ18O values do not show parallel trends and are carbon cycle and has been observed from various sections in northern
not correlated statistically; these observations support the preserva- Italy (Menegatti et al., 1998; Erba et al., 1999), Switzerland (Menegatti
tion of the original isotope values. The data clearly indicate relatively et al., 1998; Wissler et al., 2003), France (Moullade et al., 1998;
warm conditions for the late early Barremian Hauptblätterton, a Heimhofer et al., 2004), the Middle East (Vahrenkamp, 1996), the
gradual cooling throughout the early late Barremian and relative central Atlantic (Herrle et al., 2004), the Pacific (Jenkyns, 1995) and
stable cool conditions for the latest Barremian and early Aptian. elsewhere. The positive δ13C excursion in the early Aptian of the A39
section is the first record of this event in northern Europe proving that
6. Discussion the carbon shift also affected the Boreal ocean. Of interest here is the
observation that while the belemnite δ13C isotopes signal a major
6.1. Carbon isotope signature change in the carbon cycle associated with OAE 1a, the belemnite δ18O
signature does not show significant variations during this time
The general trend of increasing δ13C values during the Barremian interval. This implies that that the δ13C-excursion is not related to a
corresponds to the pattern observed in bulk carbonate samples from significant temperature change but rather reflects an independent
the Northern Tethys (Southern Alps; Weissert and Erba, 2004) and global signal related to the carbon budget of the oceans.
apparently reflects changes in the global carbon cycle. The values for
the Barremian (0‰ to 2‰ δ13C) are consistent with data gained from 6.2. Oxygen isotope values
belemnites of Hauterivian age from Speeton, UK (Price et al., 2000;
McArthur et al., 2004), showing a similar range of 0‰ to 2‰ δ13C. They From the early use of stable oxygen isotope signatures of
fit well with two belemnite populations analyzed from the early late belemnites on it has have been discussed to which extent these
Hauterivian and the mid Barremian of the Moorberg and Gott sections reflect the original temperature of the ocean water rather than a
about 40km further west (Fig. 1). A population of 121 specimens of the diagenetic overprint. It is only in recent years that interest
Hauterivian Hibolithes jaculoides from a 20cm thick interval shows a concentrated on biological factors like varying depths habitat of
range from 0‰ to 2‰ with only one specimen obtaining a value of 3‰. different genera/species, seasonal migration and ontogenetic varia-
The mid Barremian Oxyteuthis population of 64 specimens, however, tion of the life style which potentially influenced the δ18O signal.
has a much wider range from − 2‰ to 2.5‰. Our results agree with The low trace element concentrations (Mg, Mn, Fe, Sr) of the
published δ13C bulk rock data from hemipelagic sediments of the studied guards well below the critical threshold (see Section 5.3)
Vocontian Basin, which range from 1‰ to 2‰ (Föllmi et al., 2006) and suggest a preservation of the original δ18O values, except for the four
also indicate a slight trend to higher values throughout the Barremian. specimens excluded in the discussion. The homogeneous, very fine
The δ13C data of the lower part of the A39 section (0m to 30m) show grained mudstone sequence impermeable for a significant pore water
considerable fluctuations of specimens found only 20cm apart by up flow does not support any diagenetic overprint by meteoric waters.
to 2‰ (Fig. 7; Appendix B). The relative negative values of the This is supported by the aragonitic preservation of the ammonite
belemnites from the Hauptblätterton and clay–Blätterton alternation shells. The belemnite guards do not show any signs of epi- or
may reflect a regional signal of the semi-restricted Lower Saxony endobenthic life indicating anoxic conditions of bottom waters and
Basin. The high amount of terrigenous organic matter of the laminites, the absence of early diagenetic overprint due to burrowing.
which is documented both by microapalaeontological evidence The belemnite species analyzed are, with the exception of four
(spores, pollen; see Section 2.2) and macrofloral findings, imply an specimens of H. minutus and five specimens of N. ewaldi, all included in
intensive freshwater input. Periodic changes in the freshwater input the Praeoxyteuthis–Aulacoteuthis–Oxyteuthis lineage. From this belem-
may therefore explain the relative rapid shifts in the isotope signature nite stock consecutive species evolved forming a phylogenetic lineage.
of the belemnites from − 0.26‰ to 1.6‰ within one 10cm thick This evolutionary pattern may support the idea of a consistent life style
horizon. In analogy to their recent relatives, the Sepioidea and where the 11 individual species of this lineage occupied the same niche
Teuthoidea, which are the fastest growing marine invertebrates, over the ~ 6 my covering the Barremian. Since the δ18O values of the co-
belemnites may be viewed as very fast growing with a life time of occurring H. minutus correspond to those of the oxyteuthid lineage a
1year to 4years (Mutterlose, personal communication). The rapid similar depth habitat may be derived for H. minutus from these data,
ontogenetic growth of coleoids may thus explain the changes of the living in subsurface waters of perhaps 50m to 150m (see Section 5.1).
δ13C values as a result of periodic changes of freshwater influx in a The oxyteuthid belemnites discussed here are restricted to the
semi-restricted system. More stable marine conditions prevailed in Lower Saxony Basin and the adjoining North Sea; they are unknown
the upper part of the Barremian (30m to 90m). from the Tethys. The Barremian belemnite assemblages from other
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 341

Boreal areas (Greenland, Siberia), which are palaeobiogeographically belemnite faunas investigated because they are endemic to the
related to our study area, are different and only distinctly related. southern proto North Sea. Therefore we argue that the δ18O values
These biogeographic patterns exclude a major migration of the reflect the temperature conditions of a limited region. The isotope

Fig. 8. Model showing the depositional environment in the Barremian.


342 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

patterns observed can therefore not be explained by major seasonal Simultaneously the primary productivity increased during these phases.
changes or belemnite migration from one basin to another, but This in combination with generally warm conditions of 20°C to 25°C of
seasonal migration within the proto North Sea seems possible. the lower water column according to the δ18O data (SMOW = 0) and 26°C
In the A39 section the δ18O values increase from rather negative to 30°C following the TEX86 data (Mutterlose, personal communication)
(warm) values (~ − 2.0‰) in the late early Barremian (Hauptblätterton) caused a stable stratification preventing any oxygenation of the bottom
to relative positive (cool) data (~ 0‰) in the late Barremian and early waters in this quasi lagoonal system. The δ13C values vary considerably
Aptian. Some of the data show quite a considerable variation of up to from 0‰ to 2‰ in the lower and mid Barremian, less so in the upper
~ 1.0‰ in belemnite guards collected close to one another. This Barremian, where they range from 1‰ to 2‰. In general the Barremian
observation is supported by findings from a 56m thick mudstone δ13C values are significantly lower than those of the early Aptian
sequence of the Gott section 40km west of the A39 section (Fig. 1). excursion, the latter reflecting the accelerated transfer of light 12C from
There the isotope data obtained from 22 specimens, collected from one the oceans to the sedimentary carbon reservoir.
20cm thick horizon of the late early Barremian Hauptblätterton The discussed Barremian cooling trend may help in explaining the
(Mutterlose, personal communication), vary by 1.2‰ (= 4°C to 5°C) thickness variation, as well as the stratigraphic and spatial patterns of
from − 1.28‰ (21.5°C; 17.2°C) to − 2.42‰ (26.8°C; 21.5°C). This variation the laminites. The 5m to 6m thick Hauptblätterton of the early
may reflect short term climatic variations. These findings shed, Barremian, which is present throughout the Lower Saxony Basin, goes
however, doubt on the usage of individual δ18O data for palaeotem- along with a warm and stable period, allowing for a stratification of the
perature interpretations; they rather suggest that general trends based water column over a relative long time period. This warmest phase of the
on more than two or three data points in a given bio-zone should be Barremian marks the peak of laminite sedimentation, which occurred
used for palaeoclimatic interpretations. Despite these variations by both at the basin margin and in the basin centre (Fig. 8). The subsequent
~ 1.0‰ (= 4°C) for different specimens of one species from one horizon cooling trend throughout the late Barremian–early Aptian caused a
our data clearly indicate a cooling trend from a warm Hauptblätterton change towards a more instable stratification of the water column. This
period to a cool late Barremian–early Aptian. This is consistent with the in turn caused a more regular break down of the stratification resulting
temperature trends derived from the isotopic composition of fish teeth in much thinner laminates interbedded with dark clays. The lithological
and rudist bivalves (Pucéat et al., 2003; Steuber et al., 2005) and changes from laminite to clay mirror slight variations in the oxygenation
appears to reflect the global evolution of palaeoclimate. of the bottom waters from anoxic to suboxic, reflecting shifts from
Micropalaeontological observations of diversity patterns of calci- warm-humid (increased run-off, reduced surface water salinity,
spheres (Keupp and Mutterlose, 1994) from late Barremian–early Aptian increased productivity) to more cool-arid conditions (reduced run-off,
mudstones of two cores (Hoheneggelsen KB 40, KB 50; Fig. 1) approx. instable water stratification, normal productivity). The overall restricted
25km west of the A39 section indicate cooler water conditions for the conditions in this marginal basin, however, did not allow a major
clay–Blätterton alternation up to the base of the Fischschiefer. The exchange of water masses through Barremian–early Aptian times.
calcispheres are characterized by orthopithonelloids (25% to 60%) and
obliquipithonelloid (40% to 75%) cysts. The former is outer shelf dwellers 7. Conclusions
indicative for an open marine setting, the latter preferred neritic inner
shelf environments (Keupp, 1991). Within the obliquipithonelloids A continuous, 103m thick mudstone sequence of Barremian–early
changes of the abundance of coarsely and finely-crystalline morpho- Aptian age, recently exposed in northern Germany, has been logged and
types reflect the availability of calcium carbonate. The clay–Blätterton analyzed with respect to its macrofaunal content and its geochemistry
alternation with its general low carbonate content is clearly dominated (δ13C, δ18O). The palaeontological data suggest anoxic (Blätterton)–
by coarsely-crystalline, single layered morphotypes. These coarse- suboxic conditions (dark clays) of bottom waters in a semi-restricted
crystalline morphotypes make 80% to 90% of the late Barremian epicontinental sea, characterized by often endemic marine floras and
obliquipithonelloid assemblages, with fine crystalline forms becoming faunas. Short termed variations of run-off caused changes of surface
dominant in the Fischschiefer only (Keupp and Mutterlose, 1994). The water salinity and primary productivity with the laminites indicating a
two different morphotypes (fine–coarse) are thought to be phenotypical stronger input of terrestrial organic matter. This implies that the anoxic
reactions depending on primary fluctuations of the carbonate content of conditions which caused the deposition of the laminites went along
the ancient ocean water. Changes from fine–coarse and vice versa are with a periodically stable halocline which in turn resulted from
linked to changes of sea-surface temperatures, which in turn are seasonally increased run-off from the nearby hinterland.
correlated with the solubility of CaCO3. The coarse-crystalline morpho- Short termed variations of the δ13C signal from belemnites of up to
types with their single walls indicate therefore cooler conditions, while a 2‰ are explained by periodical variations of the influx of dissolved
dominance of finely-crystalline morphotypes with double layered walls inorganic carbon with low δ13C values from the nearby hinterland.
suggests a mineralization under warmer conditions. A trend towards The distinctive positive excursion of ~ 4‰ directly above the early
assemblages impoverished in calcareous phytoplankton is also reflected Aptian Fischschiefer correlates well with the Tethyan positive δ13C
by calcareous nannofossils. While the late early Barremian (Haupt- excursion of late early Aptian age above the Oceanic Event 1a. Also, the
blätterton) shows rich and diverse assemblages, diversity and abun- general trend of increasing δ13C values during the Barremian corre-
dance decrease throughout the clay–Blätterton alternation only to attain sponds to data from Tethyan deposits so that the data from the Boreal
high values in the Fischschiefer again. Realm — despite its restricted palaeoceanographic conditions — largely
reflect global perturbations in the carbon cycle.
6.3. Palaeoceanographic implications The high resolution δ18O record gained from belemnite guards of the
genera Praeoxyteuthis, Aulacoteuthis and Oxyteuthis throughout a
The palaeogeographic setting of the Lower Saxony Basin was quite complete sequence of Barremian–early Aptian mudstones clearly
similar to that of the nowadays North Sea. Both reflect shallow reflects negative (= warm) values (~ − 2‰) for the early Barremian.
epicontinental seas with water depths probably not exceeding 200m, The values increase by approximately 2‰ to values around 0‰
being surrounded by landmasses considerably larger than these indicating a temperature decrease by ~ 8°C. This late Barremian–early
marginal seas. The palynological observations of periodically reduced Aptian cooling has been observed in Tethyan sections as well, and is
surface water salinity go along with changes in the run-off from the vast consequently considered to reflect a global trend.
hinterland. During periods of increased run-off, as documented by high The late Barremian–early Aptian cooling goes along with a shift
Prasinophyta concentrations (Pterospermella, Leiosphaera) a stable from thick, up to 6.5m thick laminites (Hauptblätterton) in the early
halocline with slightly reduced surface water salinities was established. Barremian to thin 10cm to 20cm thick laminites (Blätterton) in the late
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 343

Barremian. During the early Barremian Hauptblätterton very warm (continued)A (continued)
Appendix
conditions caused a long lasting (N 500kyrs) stable water stratification, Sample no Lithology CaCO3 [%] TOC [%]
while the cooler temperatures of the late Barremian reduced the 146/2 Clay 1.89
stability of the halocline resulting in rather short termed changes of 146/3 Clay 0.00 2.14
thin laminites and dark clays. These observations suggest a tempera- 146/4 Clay 5.66
146/5 Clay 5.66 2.80
ture controlled peak of anoxic conditions in the early Barremian 148/1 Clay 8.30 1.31
(Hauptblätterton) and subsequently short termed changes from 148/2 Clay 3.77
anoxic to suboxic conditions in the late Barremian and early Aptian. 148/3 Clay 15.09
Our findings also suggest that individual belemnite δ18O data 148/4 Clay 7.92 1.13
150/1 Clay 9.43 1.03
should not be used for palaeotemperature interpretations, since
150/2 Clay 1.89 1.09
different specimens from a single layer show a δ18O variation of up to 152/1 Clay 3.77 2.01
~ 1.0‰ (~ 4°C). For interpretations of general palaeoclimatic trends 155/1 Clay 1.89 1.98
more than two or three data points in a given bio-zone should be 155/2 Clay 4.15 1.79
therefore used in future studies. Secondly, palaeobiological (way of 157/1 Clay 1.89 2.62
157/2 Clay 2.26
life) and taxonomical aspects need to be accounted for. We supply 157/3 Clay 0.94 1.25
evidence that the oxyteuthid belemnites were nektonic, active 158/1 Concretion 6.42 6.68
swimmers living in water depths of 50m to 150m. 159/1 Clay 1.89 2.05
159/2 Clay 0.00
159/3 Clay 0.00 3.77
Acknowledgments
159/4 Clay 0.00
159/5 Clay 0.00 3.64
We are grateful to K. Wiedenroth for his support during fieldwork. 161/1 Clay 0.00 2.00
U. Heimhofer stimulated our discussions with valuable comments. 161/2 Clay 0.00
Constructive reviews by K. Föllmi and an anonymous reviewer 161/3 Clay 0.00 1.95
161/4 Clay 0.00
improved the manuscript substantially. Financial support by DFG 161/5 Clay 0.00 3.01
(MU 667/28-1) is acknowledged. 161/6 Clay 0.00
161/7 Clay 11.32
Appendix A 161/8 Clay 7.17
161/9 Clay 7.17
161/10 Clay 11.70 6.19
List of bulk rock samples, TOC- and CaCO3 data from the A39 section 162/1 Concretion 35.85 7.00
Sample no Lithology CaCO3 [%] TOC [%] 163/1 Clay 0.00 4.00
163/2 Clay 4.53
99/7 Clay 18.87 0.84
163/3 Clay 0.00 3.15
99/6 Clay 26.42
163/4 Clay 0.00
99/5 Clay 14.91 2.47
163/5 Clay 0.00 2.52
99/4 Clay 14.15
165/1 Clay 12.08 6.88
99/3 Clay 11.32 0.07
165/2 Clay 0.00
99/2 Clay 6.60 1.29
165/3 Clay 7.92
99/1 Clay 42.08 4.89
165/4 Clay 26.42
100/1 Laminite; HBt. 7.17
165/5 Clay 0.00 0.41
100/2 Laminite; HBt. 18.87 4.29
165/6 Clay 0.00
100/3 Laminite; HBt. 26.04
165/7 Clay 0.00
100/3b Laminite; HBt. 10.57
165/8 Clay 15.47
100/4 Laminite; HBt. 12.08 1.54
165/9 Clay 0.00 1.46
100/5 Laminite; HBt. 28.30
167/1 Clay 3.77 1.29
100/6 Laminite; HBt. 26.42 3.87
169/1 Clay 2.83 0.84
100/7 Laminite; HBt. 24.60
169/2 Clay 1.89
100/8 Laminite; HBt. 24.53
169/3 Clay 7.55 0.81
100/9 Laminite; HBt. 17.92 3.23
171/1 Clay 8.30 2.02
100/10 Laminite; HBt. 15.47
172/1 Fischschiefer I 25.66 6.12
100/11 Laminite; HBt. 13.21
173/1 Clay 5.28 0.49
100/12 Laminite; HBt. 23.58 3.78
173/2 Clay 4.91 0.63
100/13 Laminite; HBt. 9.43 0.59
174/1 Fischschiefer II 30.94 6.03
101/1 Clay 9.43 4.41
174/2 Fischschiefer II 54.72 4.60
102/1 Laminite; Blätterton 10.38 1.78
174/3 Fischschiefer II 56.60 3.38
103/1 Clay 15.09
174/4 Fischschiefer II 58.49 2.63
105/1 Clay 9.43 0.70
175/1 Concretions 75.47 0.27
106/1 Clay 25.00
176/1 Marl 62.26 0.59
107/1 Laminite; Blätterton 15.09 5.51
176/2 Marl 11.70
108/1 Clay 15.09 1.81
176/3 Marl 42.83 0.80
109/1 Laminite; Blätterton 15.09 2.75
176/4 Clay 10.57
109/2 Laminite; Blätterton 12.64 1.42
176/5 Clay 9.06 0.16
112/1 Clay 15.09 2.31
176/6 Clay 7.17
117/1 Laminite; Blätterton 15.09 4.94
176/7 Clay 7.55 0.46
124/1 Laminite; Blätterton 7.55 3.61
177/1 Tuff 5.28 0.34
125/1 Clay 4.91 2.44
178/1 Clay 5.66 0.22
127/1 Clay 4.15 2.28
179/1 Tuff 0.00 0.03
132/1 Laminite; Blätterton 0.00 5.01
180/1 Clay 6.98 0.99
134/1 Laminite; Blätterton 0.00 6.06
181/1 Tuff 0.38 0.28
138/1 Laminite; Blätterton 0.00 6.23
182/1 Clay 4.15 0.28
141/1 Clay 0.00 5.20
183/1 Tuff 3.77 0.28
145/1 Clay 3.77 1.77
184/1 Clay 4.15 0.36
145/2 Clay 4.34 1.68
184/2 Clay 6.23 0.35
146/1 Clay 19.06 4.23
185/1 Marl 3.77 0.12
344 J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345

Appendix B (continued)B (continued)


Appendix
T [°C]
Geochemical data from Barremian belemnites, investigated for m Sample belemnite Fe Mg Mn Sr δ13C δ180 δw = δw = δw =
stable carbon, oxygen isotopes (δ13C, δ18O) and trace elements. above no. species [ppm] [ppm] [ppm] [ppm] 0 −0.5 −1
Palaeotemperatures reconstructed from δ18O (T°C). The samples base
marked with an asterisk show high trace element concentrations 22.4 ⁎130 O. germanica 182 1436 107 1647 1.25 −0.38 18 16 13
and have not been considered in Fig. 7. [1]
26.1 138/1 O. germanica 105 1389 34 1486 1.98 −0.39 18 16 14
26.2 ⁎140 O. germanica 237 2821 105 1803 1.81 −0.1 16 14 12
T [°C] base
31 145/1 O. 9 1449 2 1725 1.14 −0.29 17 15 13
m Sample belemnite Fe Mg Mn Sr δ13C δ180 δw = δw = δw =
−0.5 −1 brunsvicensis
above no. species [ppm] [ppm] [ppm] [ppm] 0
31.3 145/2 O. 227 2109 33 1788 1.09 −0.81 19 17 15
base
brunsvicensis
3.8 ⁎99/3 P. pugio 380 4898 49 2074 1.19 0.05 16 14 12 33.5 3.5m O. 76 1478 12 1675 1.89 −0.04 16 14 12
4.7 1.3m Hibolites? 48 1240 4 1568 1.36 −0.15 17 15 13 above brunsvicensis
below top
top 99 144
5.2 99/2 P. pugio 7 2479 10 2238 −0.04 −0.2 17 15 13 34.1 146/1 Oxyteuthis 119 1673 47 1573 1.95 −0.08 16 14 12
5.2 0.8m P. pugio 105 1487 8 1630 1.8 0.42 14 12 10 sp.
below 36.8 146/2 O. germanica 105 1658 34 1547 2.65 − 0.5 18 16 14
base 37 146/3 O. germanica 291 1104 48 1457 3.43 −0.57 18 16 14
100 37.7 146/4 O. germanica 131 1490 17 1777 1.5 −0.78 19 17 15
5.8 0.2m A. 138 1392 4 1752 1.1 −0.2 17 15 13 38.2 146/5 O. depressa 83 2488 3 1734 1.19 −0.09 16 14 12
below speetonensis 41 148/1 O. depressa 198 1318 34 1466 1.87 −0.08 16 14 12
base 42.3 148/2 O. germanica 99 2271 1 1491 0.52 −0.96 20 18 16
100 42.6 148/3 O. germanica 197 2712 16 1839 0.54 −0.44 18 16 14
5.9 99/1 P. pugio 17 1383 3 1874 0.36 −0.41 18 16 14 43.7 148/4 O. germanica 108 2156 3 1450 1.13 −1.1 21 19 16
7.5 100/3b A. ernsti 195 1395 19 1740 1.57 −1.16 21 19 17 46 150/1 O. depressa 186 2365 19 1824 1.66 −1.09 21 18 16
7.7 100/4 A. compressa 21 2845 2 2036 0.84 −2.03 25 23 20 47.8 2.0m O. depressa 99 1269 1 1446 1.93 −1.28 21 19 17
11.2 100/12 A. descendens 33 2057 1 1877 0.69 −2.01 25 23 20 above
12.2 100/14 A. compressa 92 1494 4 1904 1.51 −1.79 24 22 19 top
13.7 1.2m Aulacoteuthis 24 1521 4 1626 1.6 −0.77 19 17 15 149
above sp. 48.7 150/2 O. germanica 81 1807 1 1465 2.2 −0.31 17 15 13
top 50.2 152/1 O. depressa 149 1632 3 1595 1.6 −0.07 16 14 12
100 55.9 155/2 O. depressa 198 1856 3 1404 1.56 −0.18 17 15 13
14.3 103/1 O. 171 885 1 1064 1.39 −0.83 20 17 15 59.8 157/1 H. minutus 215 1652 3 1411 1.37 0.22 15 13 11
brunsvicensis 60.6 157/2 H. minutus 254 1899 2 1358 2.12 0.05 16 14 12
16.5 108/1 O. 152 1684 4 1554 1.07 −0.84 20 17 15 61.5 157/3 O. depressa 92 1469 1 1546 1.73 −0.15 17 15 13
brunsvicensis 61.8 158/1 H. minutus 99 1274 3 1276 1.62 0.22 15 13 11
17.4 0.1m O. 188 1623 21 1635 0.14 −1.07 21 18 16 74.8 0.1m O. depressa 95 913 33 1336 2.03 −0.39 18 16 14
below brunsvicensis below
top 110 base
17.5 ⁎109/2 Oxyteuthis sp. 3343 3331 135 1297 −2.53 −3.17 30 28 26 162
18.3 112 O. 232 2189 56 1566 0.53 −0.07 16 14 12 75.9 163 O. depressa 149 1652 5 1436 1.12 −0.55 18 16 14
base brunsvicensis mid
18.4 112/1 O. 328 2441 12 1592 0.35 −0.63 19 17 14 76.8 163/2 O. depressa 119 1024 2 1366 0.91 −0.11 16 14 12
brunsvicensis 77.6 163/3 O. depressa 382 1344 31 1158 1.53 0.21 15 13 11
19.4 115 O. 173 1671 27 1878 1.46 −0.94 20 18 16 83.9 165/6 O. depressa 120 1545 6 1224 2.16 0.19 15 13 11
mid brunsvicensis 84.6 2.3m O. depressa 107 968 6 1225 2.2 0.21 15 13 11
20.1 118 O. 97 1826 16 2109 0.29 −1.25 21 19 17 below
mid brunsvicensis base
20.2 120 O. 79 1368 10 1764 0.43 −0.94 20 18 16 166
base brunsvicensis 94.6 0.1m N. ewaldi 179 1832 23 1284 3.61 −0.01 16 14 12
20.3 121 O. 318 1601 26 1653 1.53 −0.17 17 15 13 above
mid brunsvicensis top 175
21.2 124 O. 31 1536 4 1627 1.62 −0.4 18 16 14 95.3 176/3 N. ewaldi 180 1809 28 1404 4.02 0.31 15 13 11
mid brunsvicensis 96.9 176/6 N. ewaldi 74 1461 8 1366 4.12 −0.35 17 15 13
21.2 124 O. 38 1598 9 1723 1.59 −0.39 18 16 14 97.7 176/7 N. ewaldi 134 2841 6 1329 1.75 0.15 15 13 11
brunsvicensis
21.2 124/1 O. 172 2094 59 2088 −0.26 −0.79 19 17 15
brunsvicensis References
21.4 125/1 O. 86 1576 20 1603 1.47 −0.36 18 15 13
brunsvicensis Anderson, T.F., Arthur, M.A., 1983. Stable isotopes of oxygen and carbon and their
21.5 126 O. germanica 392 1801 23 1731 1.23 −0.58 18 16 14 application to sedimentologic and palaeoenvironmental problems. SEPM Short
mid Course 10, 1–151.
21.6 126 O. 37 1513 3 1655 1.81 −0.11 16 14 12 Below, R., Kirsch, K.-H., 1997. Die Kerogen-Fazies der Tonstein-Blättertonstein-
top brunsvicensis Rhythmite des Ober-Barrême - Unter-Apt im Niedersächsischen Becken (Nord-
21.6 127/1 O. 246 1506 45 1798 1.92 −0.69 19 17 15 deutschland) am Beispiel der Bohrung Hoheneggelsen KB 50. Palaeontographica B
brunvicensis 242, 1–90.
21.7 128 O. germanica 816 1899 65 1427 1.22 −1.12 21 19 16 Benesch, M., Heydemann, A., Usdowski, E., 1998. Mineralogy of the Lower Aptian
sediments from the borehole Hoheneggelsen KB 40, with special emphasis on the
mid
clay mineralogy. Cretaceous Research 19, 569–580.
21.7 128 O. germanica 777 1802 59 1520 1.69 −0.84 20 17 15
Bischoff, G., Mutterlose, J., 1998. Calcareous nannofossils of the Barremian/Aptian
top
boundary interval in NW Europe: biostratigraphic and palecologic implications of a
21.7 129 O. germanica 56 1287 7 1687 1.16 −0.46 18 16 14 high resolution study. Cretaceous Research 19, 635–661.
base Brahms, H., 1913. Die Fischfauna des Barremien bei Hildesheim. 92 pp., Diss.
22.3 130 O. germanica 500 2435 24 2435 1.83 −0.49 18 16 14 Tübingen.
mid Casey, R., 1964. A monograph of the Ammonoidea of the Lower Greensand, part V.
Palaeontographical Society Monographs 117, 289–398.
J. Mutterlose et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 330–345 345

Craig, H., Gordon, L.I., 1965. Deuterium and oxygen-18 variations in the ocean and the Bedoule (southeast France). Comptes Rendus de l'Académie des Sciences. Série 2.
marine atmosphere. Proceedings of a Conference on Stable Isotopes in Oceano- Sciences de la Terre et des Planètes 327, 693–698.
graphic Studies and Paleotemperatures, Spoleto, Italy, pp. 9–130. Müller, G., Gastner, M., 1971. The “Karbonat-Bombe”, a simple device for the
Epstein, S., Buchsbaum, R., Lowenstam, H.A., Urey, H.C., 1953. Revised carbonate–water determination of the carbonate content in sediments, soils, and other materials.
isotopic temperature scale. Bulletin of the Geological Society of America 64, 1315–1325. Neues Jahrbuch für Mineralogie - Monatshefte 10, 466–469.
Erba, E., 1994. Nannofossils and superplumes: the early Aptian “nannoconid crisis”. Mutterlose, J., 1983. Phylogenie und Biostratigraphie der Unterfamilie Oxyteuthinae
Paleoceanography 9, 483–501. (Belemnitida) aus dem Barreme (Unterkreide) NW - Europas. Palaeontographica A
Erba, E., Chanell, J.E.T., Claps, M., Jones, C., Larson, R., Opdyke, B., Premoli Silva, I., Riva, A., 180, 1–90.
Salvini, G., Torricelli, S., 1999. Integrated stratigraphy of the Cismon Apticore Mutterlose, J., 1990. A belemnite-scale for the Lower Cretaceous. Cretaceous Research
(southern Alps, Italy): a “reference section” for the Barremian–Aptian interval at 11, 1–15.
low latitudes. Journal of Foraminiferal Research 29, 371–391. Mutterlose, J., 1992. Migration and evolution patterns of floras and faunas in marine
Ernst, G., 1964. Ontogenie, Phylogenie und Stratigraphie der Belemnitengattung Early Cretaceous sediments of NW Europe. Palaeogeography, Palaeoclimatology,
Gonioteuthis BAYLE aus dem nordwestdeutschen Santon/Campan. Fortschritte in Palaeoecology 94, 261–282.
der Geologie von Rheinland und Westfalen 7, 113–174. Mutterlose, J., Baraboshkin, E., 2003. The taxonomy of the early Cretaceous belemnite
Föllmi, K.B., Godet, A., Bodin, S., Linder, P., 2006. Interactions between environmental genus Aulacoteuthis Stolley, 1911 (Belemnitida). Berliner Geowissenschaftliche
change and shallow water carbonate buildup along the northern Tethyan margin Abhandlungen 3, 179–187.
and their impact on the Early Cretaceous carbon isotope record. Paleoceanography Mutterlose, J., Böckel, B., 1998. The Barremian–Aptian boundary interval in NW
21. doi:10.1029/2006PA001313. Germany — A synthesis. Cretaceous Research 19, 539–568.
Habermann, A., Mutterlose, J., 1998. Early Aptian black shales from NW Germany: Mutterlose, J., Bornemann, A., 2000. Distribution and facies patterns of Lower
calcareous nannofossils and their palaeoceanographic implications. Neues Jahr- Cretaceous sediments in northern Germany — a review. Cretaceous Research 21,
buch für Geologie und Paläontologie, Abhandlungen 212 (1–3), 379–400. 733–759.
Heimhofer, U., Hochuli, P.A., Herrle, J.O., Andersen, N., Weissert, H., 2004. Absence of Mutterlose, J., Harding, I., 1987. Phytoplankton from the anoxic sediments of the
major vegetation and palaeoatmospheric pCO2 changes associated with oceanic Barremian (Lower Cretaceous) of North–West-Germany. Abhandlungen Geolo-
anoxic event 1a (Early Aptian, SE France). Earth and Planetary Science Letters 223, gische Bundes-Anstalt 39, 177–215.
303–318. Price, G.D., Ruffell, A.H., Jones, C.E., Kalin, R.M., Mutterlose, J., 2000. Isotopic evidence for
Herrle, J.O., Kössler, P., Friedrich, O., Erlenkeuser, H., Hemleben, C., 2004. High resolution temperature variation during the early Cretaceous (late Ryazanian-mid-Hauter-
carbon isotope records of the Aptian to lower Albian from SE France and the ivian). Journal of the Geological Society (London) 157, 335–343.
Mazagan Plateau (DSDP Site 545). A stratigraphic tool for paleoceanographic and Pucéat, E., Lécuyer, C., Sheppard, S.M.F., Dromart, G., Reboulet, S., Grandjean, P., 2003.
paleobiologic reconstruction. Earth and Planetary Science Letters 218, 149–161. Thermal evolution of Cretaceous Tethyan marine waters inferred from oxygen
Hild, E., Brumsack, H.J., 1998. Major and minor element geochemistry of Lower Aptian isotope composition of fish tooth enamels. Paleoceanography 18, 1029. doi:10.1029/
sediments from the NW German Basin (core Hoheneggelsen KB 40). Cretaceous 2002PA000823.
Research 19, 615–634. Raisossadat, S.N., 2004. The ammonite family Deshayesitidae in the Kopet Dagh Basin,
Hochuli, P.A., Menegatti, A.P., Weissert, H., Riva, A., Erba, E., Premoli Silva, I., 1999. north-east Iran. Cretaceous Research 25, 115–136.
Episodes of high productivity and cooling in the early Aptian Alpine Tethys. Geology Rawson, P.F., Riley, L.A., 1982. Latest Jurassic–early Cretaceous events and the “late
27, 657–660. Cimmerian unconformity” in North Sea area. American Association of Petroleum
Jenkyns, H.C., 1995. Carbon-isotope stratigraphy and palaeoceanographic significance of Geologists Bulletin 66, 2628–2648.
the lower Cretaceous shallow water carbonates of Resolution Guyot, Mid-Paci¢c Rückheim, S., Mutterlose, J., 2002. The Early Aptian radiation of planktonic foraminifera
mountains. Proceedings of the Ocean Drilling Program. Scientific Results 143, 99–104. in NW Europe: the onset of the Mid-Cretaceous plankton revolution in the Boreal
Kemper, E., 1971. Zur Gliederung und Abgrenzung des norddeutschen Aptium mit Realm. Cretaceous Research 23, 49–64.
Ammoniten. Geologisches Jahrbuch 89, 359–390. Ruffell, A., 1991. Sea-level events during the early Cretaceous in western Europe.
Kemper, E., 1979. Die Unterkreide Nordwestdeutschlands. Ein Überblick. In: Wiedmann, Cretaceous Research 12, 527–551.
J. (Ed.), Aspekte der Kreide Europas. IUGS Series A, vol. 6. Schweizerbart, Stuttgart, Schott, W., Jaritz, W., Kockel, F., Sames, C.W., Stackelberg, V., Stets, J., Stoppel, D.,
pp. 1–9. Baldschuhn, R., Krampke, K.D., 1967/69. Paläogeographischer Atlas der Unterkreide
Kemper, E., 1995. Die Entfaltung der Ammoniten und die Meeresverbindungen im von Nordwestdeutschland mit einer Übersichtsdarstellung des nördlichen Mitte-
borealen Unter- und Mittel-Apt. Geologisches Jahrbuch A 141, 171–199. leuropa. 306 Kt., (Hannover). Hierzu Erläuterungen zum Paläogeographischen Atlas
Keupp, H., 1991. Fossil calcareous dinoflagellate cysts. In: Riding, R. (Ed.), Calcareous der Unterkreide von NW-Deutschland. 315 S. (Hannover).
algae and stromatolites. Springer, Berlin, pp. 267–286. Shackleton, N.J., Kennett, J.P., 1975. Paleotemperature history of the Cenozoic and the
Keupp, H., Mutterlose, J., 1994. Calcareous phytoplankton from the Barremian/Aptian initiation of Antarctic glaciation: oxygen and carbon isotope analyses in DSDP sites
boundary interval from NW Germany. Cretaceous Research 15, 739–763. 277, 279, and 281. In: Kennett, J.P., Houtz, R.E. (Eds.), Initial Reports of the Deep Sea
Littke, R., Jendrzejewski, L., Lokay, P., Wang Shuangqing, Rullkötter, J., 1998. Organic Drilling Project, vol. 29, pp. 743–756.
geochemistry and depositional history of the Barremian–Aptian boundary interval Steuber, T., 1999. Isotopic and chemical intra-shell variations in low-Mg calcite of rudist
in the Lower Saxony Basin, northern Germany. Cretaceous Research 19, 581–614. bivalves (Mollusca–Hippuritacea): disequilibrium and late Cretaceous seasonality.
Marek, S., Racynska, A., 1979. Paläogeographie der Unterkreide des nordpolnischen International Journal of Earth Sciences 88, 551–570.
Beckens. In: Wiedmann, J. (Ed.), Aspekte der Kreide Europas. IUGS Series A6. Steuber, T., 2001. Strontium isotope stratigraphy of Turonian–Campanian Gosau-type
Schweizerbart, Stuttgart, pp. 447–462. rudist formations in the Northern Calcareous and Central Alps (Austria and
McArthur, J.M., Mutterlose, J., Price, G.D., Rawson, P.F., Russel, A., Thirlwall, M.F., 2004. Germany). Cretaceous Research 22, 429–441.
Belemnites of Hauterivian and Barremian (early Cretaceous) age: Sr isotope Steuber, T., Rauch, M., Masse, J.-P., Graaf, J., Malkoc, M., 2005. Low-latitude seasonality of
stratigraphy (87Sr/86Sr) and kinetic control of stable isotopic and trace element Cretaceous temperatures in warm and cold episodes. Nature 437, 1341–1344.
composition δ13C, δ18O, Na, Sr, Mg). Palaeogeography, Palaeoclimatology, Palaeoe- Taverne, L., Ross, P.H., 1973. Fischreste aus dem Töck (Unter-Aptien) von Helgoland.
cology 202, 253–272. Meyniana 23, 99–111.
McArthur, J.M., Janssen, N.M.M., Reboulet, S., Leng, M.J., Thirlwall, M.F., van de Thomsen, E., 1989. Seasonal variation in Boreal Early Cretaceous Calcareous Nanno-
Schotbrugge, B., 2007. Palaeotemperatures, polar ice-volume, and isotope strati- fossils. Marine Micropaleontology 15, 123–152.
graphy (Mg/Ca, δ18O, δ13C, 87Sr/86Sr): the Early Cretaceous (Berriasian, Valanginian, Vahrenkamp, V.C., 1996. Carbon isotope stratigraphy of the upper Kharaib and Shuaiba
Hauterivian). Palaeogeography, Palaeoclimatology, Palaeoecology 248, 391–430. Formations: implications for the Early Cretaceous evolution of the Arabian Gulf
Menegatti, A.P., Weissert, H., Brown, R.S., Tyson, R.V., Farrimond, P., Strasser, A., Caron, region. AAPG Bulletin 80, 647–662.
M., 1998. High-resolution δ13C stratigraphy through the early Aptian “Livello Selli” Weissert, H., Erba, E., 2004. Volcanism, CO2 and palaeoclimate: a Late Jurassic–Early
of the Alpine Tethys. Paleoceanography 13, 530–545. Cretaceous carbon and oxygen isotope record. Journal of the Geological Society
Michael, E., 1967. Die Mikrofauna des nw-deutschen Barrême. Teil 1 Die Foraminiferen (London) 161, 695–702.
des nw-deutschen Barrême. Palaeontographica A12, 1–176. Weissert, H., Lini, A., Föllmi, K.B., Kuhn, O., 1998. Correlation of Early Cretaceous carbon
Michael, E., 1974. Zur Palökologie und Faunenführung des norddeutschen Unterkreide- isotope stratigraphy and platform drowning events: a possible link? Palaeogeo-
Meeres. Geologisches Jahrbuch A 19, 1–68. graphy, Palaeoclimatology, Palaeoecology 137, 189–203.
Michael, E., 1979. Mediterrane Fauneneinflüsse in den borealen Unterkreide-Becken Wissler, L., Funk, H., Weissert, H., 2003. Response of Early Cretaceous carbonate
Europas, besonders Nordwestdeutschlands. In: Wiedmann, J. (Ed.), Aspekte der platforms to changes in atmospheric carbon dioxide levels. Palaeogeography,
Kreide Europas. IUGS Series A, vol. 6. Schweizerbart, Stuttgart, pp. 305–321. Palaeoclimatology, Palaeoecology 200, 187–205.
Moullade, M., Kuhnt, W., Bergen, J.A., Masse, J.P., Tronchetti, G., 1998. Correlation of
biostratigraphic and stable isotope events in the Aptian historical stratotype of La
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / p a l a e o

Late Pliensbachian–Early Toarcian (Early Jurassic) environmental changes in an


epicontinental basin of NW Europe (Causses area, central France): A
micropaleontological and geochemical approach
Samuel Mailliot a, Emanuela Mattioli a,⁎, Annachiara Bartolini b, François Baudin c,
Bernard Pittet a, Jean Guex d
a
UMR 5125 PEPS, CNRS, France; Université Lyon 1, Campus de la DOUA, Bâtiment Géode, 69622 Villeurbanne Cedex, France
b
UMR 5143, CNRS, France; Muséum National d'Histoire Naturelle, USM 203, 8 rue Buffon CP 38 75231 Paris cedex 05, France
c
Université Pierre et Marie Curie-Paris6 and UMR 7072 Tectonique, CNRS, CP 117, 4 place Jussieu 75252 Paris Cedex 05, France
d
Institut de Géologie et Paléontologie, Université de Lausanne, Anthropole, CH 1015 Lausanne, Switzerland

A R T I C L E I N F O A B S T R A C T

Article history: We present an integrated work based on calcareous nannofossil and benthic foraminiferal assemblages, and
Received 30 November 2007 geochemical analyses of two Upper Pliensbachian–Lower Toarcian sections located in the central-South
Received in revised form 28 May 2008 France. The studied sections, Tournadous and Saint-Paul-des-Fonts, represent the proximal and the distal
Accepted 28 May 2008
part, respectively, of the Jurassic Causses Basin, one of the small, partly enclosed basins belonging to the
epicontinental shelf of the NW Tethys. At the transition from Late Pliensbachian to Early Toarcian, the
Keywords:
Causses Basin recorded an emersion in response to the global sea-level fall. Our data indicate severe
Calcareous nannofossils
Benthic foraminifers
environmental conditions of marine waters, including salinity decrease and anoxia development, occurring
Organic matter in the Early Toarcian. The acme of this deterioration coincides with the Early Toarcian Anoxic Event (T-OAE)
Carbon and oxygen stable isotopes but, due to the restricted nature of the basin, anoxia persisted until the end of the Early Toarcian, mainly in
Pliensbachian/Toarcian the deeper parts of the basin. The micronutrients and organic organic-matter fluxes were probably high
Toarcian Anoxic Event (T-OAE) during the entire studied time interval, as shown by nannofossil and foraminiferal assemblages. However,
nannoplankton production drastically decreased during the T-OAE, as demonstrated by very low nannofossil
fluxes, and only taxa tolerant to low-saline surface waters could thrive. At the same time, benthic
foraminifers temporarily disappeared in response to sea-bottom anoxia. Our study demonstrates that
environmental changes related to the T-OAE are well-recorded even in small, partly enclosed basins of NW
Europe, like the Causses Basin. Within this area, the effects of global changes, like sea sea-level and
temperature fluctuations, are modulated by local conditions mainly controlled by the morphology of the
basin.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction Jenkyns et al., 2001; Röhl et al., 2001; Jenkyns et al., 2002; van Breugel
et al., 2006; Emmanuel et al., 2006; Suan et al., 2008a), and
The Early Toarcian is marked by major environmental perturba- continental material (fossil wood, Hesselbo et al., 2000, 2007)
tions. In many epicontinental basins, organic-rich sediments are recorded in the Exaratum Ammonite Subzone that reflects major
recorded that are interpreted as deposited under anoxic conditions. perturbations in the global carbon cycle. In the rest of this account, we
The widespread record of such deposits led Jenkyns (1988) to refer to as T-OAE the stratigraphic interval corresponding to enrich-
interpret the Early Toarcian anoxia as a global event (T-OAE). Anoxia ment in organic matter of the sediments and to the carbon isotope
was temporarily spreading into the photic zone, as indicated by the negative excursion. This interval, according to various authors,
presence of biomarkers of green sulphur bacteria in some black shales corresponds to the Exaratum Ammonite Subzone (Jenkyns and
(Schouten et al., 2000; Schwark and Frimmel, 2004; Pancost et al., Clayton, 1997; Hesselbo et al., 2000; Bailey et al., 2003) and to the
2004; van Breugel et al., 2006). The organic-rich sediments are C. superbus nannofossil zone (Mattioli et al., 2004; Mailliot et al.,
associated to a pronounced negative excursion in δ13C of both marine 2006).
(bulk carbonate, brachiopod calcite, organic matter, and specific Faunal extinctions in the marine realm occurred concomitant with
organic compounds; Jenkyns and Clayton, 1986; Schouten et al., 2000; the T-OAE, but also at the Pliensbachian/Toarcian boundary (Little and
Benton, 1995; Harries and Little, 1999; Pálfy and Smith, 2000; Wignall,
⁎ Corresponding author. 2001; Cecca and Macchioni, 2004; Wignall et al., 2005). The T-OAE
E-mail address: emanuela.mattioli@univ-lyon1.fr (E. Mattioli). may represent the acme of a longer-term environmental crisis

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.05.014
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 347

(Wignall et al., 2006). An important cooling of seawaters occurred in South France). This basin, that was part of the NW Europe epi-
the Late Pliensbachian, as inferred by oxygen isotope ratio and Mg/Ca continental shelf of the western Tethys, recorded the sea-level
measured on belemnite rostra (McArthur et al., 2000; Bailey et al., changes occurring in the studied interval (Trümpy, 1983). Sea-level
2003; Rosales et al., 2004; van de Schootbrugge et al., 2005a,b), and variations were evaluated on the basis of the analysis of sedimentary
δ18O of brachiopod shells (Suan et al., 2008a). A significant marine structures observed in the two studied sections. This integrated
regression and emersions are recorded between the end of the approach enabled us to reconstruct the evolution of the entire water
Pliensbachian and the base of the Toarcian in various western Tethyan column conditions in terms of productivity and oxygenation across
settings. This record has been interpreted as evidence for glacio- the T-OAE.
eustatism (Guex et al., 2001; Morard et al., 2003).
This study aims to reconstruct the environmental changes 2. Geological overview
occurred in a time interval spanning the Late Pliensbachian and the
Early Toarcian. We present here an integrated study on planktonic The study area is located in the central-South France (Fig. 1). The
micro-organisms (calcareous nannofossils), benthic fauna (foramini- Causses Basin was a small, partly enclosed, intracratonic basin that
fers), organic matter and stable isotopes from two sections, Tourna- was part of the epicontinental seas at the West of the Tethys Ocean, at
dous and Saint-Paul-des-Fonts, located in the Causses Basin (central- a palaeolatitude comprised between 25 and 30° N (Fig. 1a). It was

Fig. 1. (a) Palaeogeographic map of the western Tethys (European realm) re-drawn after Bassoulet et al. (1993). The study area (central-South France) was located in the wide
epicontinental shelf of the NW Tethyan margin. (b) Simplified geologic map of the Causses Basin (after Trümpy, 1983) showing the location of the two studied sections: Tournadous
(TND) and Saint-Paul-des-Fonts (SPF).
348 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

located between the southern part of the French Massif Central and wise temperature pyrolysis (300–600 °C), and is expressed in hydro-
the northern part of the Montagne Noire (Fig. 1b). The basin was carbons (HC) per gram of dry sediment. Although the type of organic
bounded by Hercynian crystalline rocks, belonging to different units of matter is usually defined by mean of elemental analysis, the HI
the Massif Central, namely the Rouergue to the West, the Cevennes to approximates the H/C atomic ratio, which is determinant for organic-
the East and Aubrac to the North (Trümpy, 1983). The morphology of matter characterization (Tissot and Welte, 1984). The hydrogen index
the basin was largely controlled by the late Hercynian structural (HI) is the normalized S2 value (HI = S2 / TOC ⁎ 100) expressed in
evolution. In the Early Jurassic, an extensional tectonic phase milligram HC per gram of TOC. It is characteristic of the type of
reactivated some NNE–SSW normal faults (Arthaud and Matte, organic matter, terrestrial or marine (Tissot and Welte, 1984; Espitalié,
1975). Subsidence was more pronounced in the central part of the 1993). Tmax is the temperature at maximum pyrolytic hydrocarbon
basin. The thickness of the Lower Jurassic succession strongly varies generation expressed in °C. For a given organic-matter type, Tmax
from the basin margins to the depocenter, for instance the Toarcian varies as a function of the natural thermal maturity of the organic
deposits vary from 15 cm to 15 m in the different localities (Morard, matter.
2004). The sediments deposited on the margins of the basin are also The origin of the organic matter can be evaluated using the
characterized by the presence of repeated hiatuses (Trümpy, 1983). relationship between HI vs. Tmax (Espitalié et al., 1985). According to
Two sections have been analysed in this study (Fig. 1b), namely these authors, the organic matter can be classified in four types. Types
Tournadous (TND, very proximal) and Saint-Paul-des-Fonts (SPF, I and II correspond to fresh organic matter of lacustrine and marine
more distal). The study interval corresponds to the Lower Pliensba- origin, respectively. Type III is derived from terrestrial higher plants.
chian Marnes de Villeneuve Fm. (Trümpy, 1983) and the Lower Type IV represents highly oxidized oxidised organic matter (Peters,
Toarcian Schistes carton Fm. (Fig. 2). The Marnes de Villeneuve Fm. is 1986). The HI vs. Tmax plots of Tournadous and Saint-Paul-des-Fonts
represented by dark marls with rare, interbedded limestone beds are shown in Fig. 5.
often bioturbated and laterally discontinuous (Trümpy, 1983), inter-
preted as diagenetic in origin (Morard, 2004). A hard-ground enriched 3.1.2. Carbon and oxygen stable isotopes
in oxidized oxidised and phosphatized material, corroded belemnite All stable isotope analyses were carried out on bulk carbonates
rostra, and fossil wood marks the Pliensbachian/Toarcian boundary. (31 samples, Tournadous section; Fig. 3). Previous works have proven
This level corresponds to a hiatus (Trümpy, 1983) and is interpreted the usefulness of bulk rock analysis in pelagic fine-grained and
here as an emersion surface. The Lower Toarcian Schistes carton show homogeneous rocks (e.g., Scholle and Arthur, 1980; Anderson and
finely laminated shales at the base (Schistes Papyracés according to Arthur, 1983). Carbonate δ13C and δ18O values were measured in a
Monestier, 1921), and contain some detritic beds, enriched in silty Finningan Mat delta-S Mass Spectrometer at the Laboratory of Isotope
material, showing Hummocky Cross Stratification (HCS). One of these Geochemistry of the University of Lausanne. Samples were measured
beds is rich in fish remains (Leptolepis bed, Trümpy, 1983; Fig. 2) and is for their carbonate δ13C and δ18O content, following the standard
present across the whole basin. Ammonites, belemnites, small procedure of Mc Crea (1950), reacted at 50 °C for half an hour. A
bivalves, and fish remains are frequent in the Schistes carton (Trümpy, fractionation factor of 1.00931 between CaCO3 and H3PO4 was applied
1983). The two sections are precisely correlated by means of (Swart et al., 1991). All δ13C and δ18O values are reported relative to
ammonite and calcareous nannofossil biostratigraphy, and lithostrati- VPDB. The standard deviation of the analyses is ±0.05‰ for δ13C,
graphy (Fig. 2). and ±0.1‰ for δ18O.

3. Materials and methods 3.2. Calcareous nannofossil

3.1. Geochemistry Sixty-nine smear slides were prepared for calcareous nannofossil
quantitative analysis (38 from SPF and 31 from TND; Fig. 2). Smear
3.1.1. Carbonate and organic-matter content slides were prepared following the method described by Geisen et al.
The samples (51 for TND and 70 for SPF; Figs. 3 and 4) were cleaned (1999). A small amount of powdered rock (~ 30 mg) was dried and
and dried in an oven at 90 °C for 24 h, before they were powdered in a mixed to water (with a basic pH, and over-saturated with respect to
rotating disc mill. Total carbon (TC, wt.%) was determined using a LECO carbonates) in a magnetic stirrer in order to obtain a homogeneous
IR 212 analyser by combustion of 100 mg of sample up to 1050 °C, suspension. The suspension was let settle on a cover slide in a set-
without previous decarbonatation; the resulting CO2 was quantified by tling device for 24 h. The cover slide was then recovered, dried and
infrared detector. Calcium carbonate content (CaCO3, wt.%) was mounted on a microscope slide. Smear slides were studied with a
determined using a calcimetric bomb by HCl digestion of 100 mg of ZEISS polarizing optical microscope at 1250×. Preservation of calcar-
sample; the resulting CO2 was determined by pressure measurement. eous nannofossils was estimated on the basis of the degree of etching
Total organic carbon content (TOC, wt.%) was calculated by determin- and overgrowth (Roth, 1984), and the fragmentation of specimens.
ing the difference between total carbon and carbonate carbon, Three preservation classes were distinguished, namely: 1) poorly
assuming that all carbonate is pure calcite and represents the unique preserved assemblages (P) with heavily etched or heavily overgrown
source of inorganic carbon. The precision for these analyses is ±0.5% specimens, commonly fragmented; 2) moderate preservation (M)
for calcium carbonate and ±0.1% for total and organic carbon. presenting slightly etched or slightly overgrown specimens; 3) good
Information on the type and thermal maturity of the organic preservation (G) when etching or overgrowth are very limited, or
matter was obtained using Rock-Eval ® pyrolysis with an Oil Show coccolith structure is pristine.
Analyser device (Vinci Technologies) under standard conditions (see Between 150 and 400 calcareous nannofossils (both coccoliths and
all details about Rock-Eval pyrolysis and parameters in Espitalié, 1993). Schizosphaerella spp., incertae sedis probably a dinoflagellate cyst,
The parameter S2 is the amount of hydrocarbon released during step- Kälin and Bernoulli, 1984) were counted in a variable number of

Fig. 2. Litho- and biostratigraphy of the studied sections. Ammonite zones found in this section are after Guex (1972), Mattei et al. (1967) and Morard (2004). The two sections are
precisely correlated on the basis of integrated ammonite and nannofossil biostratigraphy. Some marker beds (i.e., the Leptolepis bed) are also used to better correlate the two sections.
The first ammonite horizon in SPF corresponds to the Semicelatum Subzone, belonging to the first Toarcian ammonite zone; whilst at TND the first ammonite horizons indicate the
second ammonite zone of the Toarcian. Note that the FO of C. poulnabronei, C. superbus, and D. ignotus are recorded in the same sample in TND, whilst these events are observed in
different samples in SPF. Absolute abundance of calcareous nannofossils (coccoliths plus Schizosphaerella) is compared to wt.% CaCO3. Preservation state of each sample is also
displayed. Absolute abundance of Schizosphaerella is shown separately (empty dots).
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 349
350 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

fields of view. In few, very poor samples, a significant number of – morphogroup C: uniserial, elongated and cylindrical forms (Proden-
nannofossils (100) could not be counted. Absolute abundance of talina, Mesodentalina, Pseudonodosaria, Nodosaria, Marginulina);
nannofossil per gram of rock was calculated using the following – morphogroup D: flat, evolute, planispiral forms (Astacolus, Lenticulina);
formula (after Geisen et al., 1999): – morphogroup E: lenticular, strongly biconvex, involute, planispiral
foraminifers (Lenticulina);
x ¼ ðNTV Þ=ðmTFTSThÞ – morphogroup F: discoid test, proloculus followed by undivided
tubular second chamber (Spirillina);
where x is the number of calcareous nannofossils per gram of rock; – morphogroup G: trochospiral, conical forms (Epistomina, Reinholdella).
N is the number of nannofossils counted in a sample; V is the volume
of water used for the dilution in the settling device; m is the weight of The morphology of the tests of benthic foraminifers is strongly
the powder used for the suspension; F is the number of fields of view influenced by life mode (e.g., infauna vs. epifauna, free vs. fixed, etc.) and
in which calcareous nannofossils were counted; S is the surface of a by environmental conditions in their habitat (depth, light intensity,
field of view; h is the height of the water over the cover slide in the hydrodynamic energy, oxygen availability; Hallock, 1979; Corliss, 1985;
settling device. Species-specific absolute and relative abundance Bernhard, 1986; Corliss and Chen, 1988; Corliss, 1991; Bartolini et al.,
were also calculated. Coccolith relative abundance was estimated 1992; Rey et al., 1994; Jorissen, 1999). Each of these morphogroups
with respect to the number of coccoliths counted in a slide (Figs. 6 and should therefore correspond to peculiar environmental conditions (e.g.,
7). The samples in which less than 100 coccoliths could be counted sea sea-bottom oxygenation, hydrodynamic energy, etc.). For the genus
were not taken into consideration for calculation of coccolith Lenticulina, the proportion of the morphogroup E vs. the morphogroup D
percentage. (E /E +D ⁎ 100) has been calculated and reported in Fig. 8.
On the basis of morphological similarity with recent taxa, we
3.3. Benthic foraminifers hypothesized that morphogroups A, B, and C may have corresponded
to infaunal foraminifers, and morphogroups D, E, F, and G to epifaunal
Thirty-one samples were processed for foraminiferal analyses: forms (see also Corliss and Chen, 1988; Corliss, 1991). The relative
14 samples from the Tournadous section and 17 samples from the abundance of infaunal vs. epifaunal morphogroups was calculated; it
Saint-Paul-des-Fonts section. About 300 g of marl samples were was is shown in Fig. 3. However, the relationship between test morphology
submitted to a thermal cracking (the sample was put several times in a and microhabitat is complex, and all generalisations have significant
oven and in cold water, alternatively). After the drying and sieving of exceptions. Caveats in using morphogroup analysis to infer micro-
residues through 1 mm, 500 μm, 250 μm, 125 μm and 50 μm meshes, habitats and environmental parameters have been exhaustively
the foraminifers were picked up for all the fractions N125 μm under a discussed by many authors (e.g., Buzas et al., 1993; Jorissen, 1999;
binocular microscope. We however checked also the finest fraction Murray, 2001). For instance, in their statistical study, Buzas et al.
(50 μm) to check the presence/absence of foraminifera, especially in (1993) evaluated that for the taxa they considered, microhabitat
the “anoxic” levels. No foraminifers were found in the finest fraction in assignments on the basis of morphology were about 75% accurate. We
these levels. Between 100 and 650 benthic foraminifers per sample can thus argue that only major changes in percentages of the
were identified and counted in the size fraction N125 μm. The fora- morphogroups are likely significant.
miniferal genera classification is based on the work of Loeblich and
Tappan (1988). In the finest fraction (50 μm), only the presence/ 4. Results
absence of the foraminifers was checked.
In the Late Pliensbachian and Early Toarcian, foraminiferal 4.1. Geochemistry
assemblages were richer at Tournadous than at Saint-Paul-des-
Fonts. In the latter section, only five samples from the Spinatum 4.1.1. Carbonate and organic-matter content
Ammonite Zone (Late Pliensbachian) yielded foraminifers, whereas all Calcium carbonate content (wt.% CaCO3) is relatively low and stable in
the Early Toarcian samples were barren. For this reason, more detailed the samples from the Upper Pliensbachian both in the Saint-Paul-des-
analyses have been performed only at Tournadous and data from the Fonts section (on average 30%) and Tournadous (20%), with the exception
five productive samples in Saint-Paul-des-Fonts are not shown. of few values measured in nodular limestones (Fig. 2). The samples from
In order to evaluate some taphonomic processes (selective the basal Toarcian are nearly depleted in CaCO3 in both sections. A
dissolution, hydrodynamic energy, etc.), which could have affected carbonate recovery is observed from the end of the Exaratum Ammonite
foraminiferal assemblages, the relative abundance of specimens Subzone and, mainly, in the Elegans Ammonite Subzone (on average 45%).
fragmented or bearing signs of dissolution, abrasion and micro-boring With the exception of a coaly horizon at the Pliensbachian/
was estimated (Fig. 8). The variations in the relative abundance of Toarcian boundary in the Saint-Paul-des-Fonts section, in which TOC
fragmented and etched specimens have been compared to those of content attains 58% (Fig. 4), organic content of the studied samples
robust and biconvex Lenticulina. Experimental evidence shows that varies between 0 and 11% and is roughly correlated to lithological
Lenticulina, a hyaline foraminifer, is one of the most resistant taxa to changes. The Upper Pliensbachian marls usually show the lowest
dissolution (e.g., Nguyen and Speijer, 2007). The relative abundance of values (b1%), with the exception of few samples, whereas a sharp
the hyaline dominant genera has been calculated and shown in Fig. 8, increase in TOC content is observed along with the deposition of the
together with the relative abundance of porcellanaceous (mainly Schistes carton facies (8% on average).
Ophtalmidium) and siliceous agglutinated tests (mainly Glomospirella The low Tmax values (average of 425 °C and 426 °C at Saint-Paul-
and Ammodiscus). The genus richness (number of genera) for each des-Fonts and Tournadous, respectively) indicate that the organic
sample is illustrated as well (Fig. 8). matter did not experience high temperatures during burial. A quite
On the basis of the external test geometry and symmetry, benthic immature organic matter is also documented by Farrimond et al.
foraminifers were attributed to 7 morphogroups, mainly following (1989) for time-equivalent sediments from North-East France
Bernhard (1986) and Bartolini et al. (1992). (Paris Basin). As illustrated in a HI-Tmax diagram (Fig. 5), the studied
samples are immature with respect to oil generation. Consequently,
– Morphogroup A: uniserial, elongated and thin specimens belong- the type of organic matter can be estimated from pyrolysis data on the
ing to the genera Paralingulina and Ichthyolaria; basis of the Hydrogen Index (HI) parameter. According to the wide
– morphogroup B: elongated and spindle-shaped foraminifers, such range of HI-values recorded (14 to 729 mg HC/g TOC), the organic
as Eoguttulina; matter of the studied samples can be attributed to Types II and III
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364
Fig. 3. Wt% TOC, HI, and Tmax are compared to the carbon and oxygen isotope values measured in the TND section. Empty dots correspond to samples affected by a diagenetic overprint. The shaded area corresponds to the T-OAE. The relative
abundance of infaunal and epifaunal morphogroups indicates sea-bottom oxygenation.

351
352 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

Fig. 4. Wt% TOC, HI, and Tmax values measured in the SPF section. The shaded area corresponds to the T-OAE.

(Fig. 5). Type II corresponds to fresh organic matter of marine origin. In the smear slides corresponding to the samples enriched in
Type III is derived from terrestrial higher plants (Espitalié et al., 1985; organic matter, pyrite framboids are commonly recorded. In discrete
Peters, 1986). levels, associated to the framboids are small (1 µm) cubic or
HI is low in the Late Pliensbachian and, similarly to TOC, its tetrahedral crystals probably corresponding to magnetite (Fe3O4) or
values increase in both sections in the Early Toarcian (Figs. 3 and griegite (Fe3S4, precursor of pyrite, FeS2). These can be isolated or
4). In the Late Pliensbachian, average values of HI are 35 mg HC/g arranged in clusters of about twelve microcrystals.
TOC at Tournadous, and 130 mg HC/g TOC at Saint-Paul-des-Fonts.
In the Early Toarcian, HI is on average 424 mg HC/g TOC at 4.1.2. Carbon and oxygen stable isotopes measured at Tournadous
Tournadous (maximal values = 729) and 496 mg HC/g TOC at Saint- The carbon isotope values are comprised between −1‰ and +1‰
Paul-des-Fonts (maximal values = 715). Values higher than 400 mg in the Late Pliensbachian, with the exception of few, very negative
HC/g TOC are generally related to marine organic matter (Type II), values (−8 and −9‰) recorded in the nodular limestones, diagenetic in
dominated by phytoplankton and bacteria with limited proportions origin, or in the coarse detritic levels bearing HCS structures (Fig. 3). In
of terrestrial organic matter. Organic matter is of Type III these levels, carbon isotope values negatively covary with δ18O. The
(terrestrial OM) in the Late Pliensbachian of the two sections, very negative carbon isotope values of these samples, and the negative
while marine organic matter (Type II) dominates in the Early correlation between carbon and oxygen isotopes are likely the result
Toarcian (Fig. 5). of a diagenetic overprint (Raiswell, 1988; Sass et al., 1991; Patterson
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 353

Fig. 5. Type of organic matter as derived by the HI vs. pyrolysis Tmax plot. Organic matter is immature in all the studied samples. It is terrestrial in origin in the samples from the
Pliensbachian in both sections but marine in the Toarcian.

and Walter, 1994). The well-known negative excursion of carbon interval (Röhl et al., 2001). These authors interpret such negative
isotopes (Hesselbo et al., 2000), occurring in different NW Europe values in terms of important fresh-water input from rivers surround-
settings in the Exaratum Ammonite Subzone, is probably recorded in the ing the Quercy and south-west German Basins, respectively. Similarly,
Tournadous section, although in few (four) samples. In fact, at the base of the Causses Basin was bounded by emerged lands. It likely received
the Toarcian (more precisely, just above the Pliensbachian/Toarcian significant riverine inputs, mainly during the Early Toarcian when
hiatus), δ13C values decrease to ~−4‰ (Fig. 3). These negative carbon an enhanced hydrological cycle is inferred on the basis of osmium
isotope values correlate to the first increase (up to 10%) of TOC. In the isotopes (Cohen et al., 2004).
Saint-Paul-des-Fonts section, the δ13C negative excursion (~−4‰) is also If we assume that the fresh-water input produced a fairly constant
recorded in very few samples from the Exaratum Ammonite Subzone shift towards more negative oxygen values for all the samples, the
(Morard, 2004). general trend of the curve in Fig. 3 is consistent with data presented in
Interestingly, carbon isotope values from one section of the Quercy the palaeotemperature dealing with palaeotemperature fluctuations.
Basin (Penne, SW France; Emmanuel et al., 2006), which was very Indeed, various authors (McArthur et al., 2000; Röhl et al., 2001;
close to the Causses Basin, are in the same range as in the Tournadous Bailey et al., 2003; Rosales et al., 2004; Suan et al., 2008a) have
section. Emmanuel et al. (2006) also record a negative δ13C excursion inferred a temperature minimum in the Late Pliensbachian, followed
of about −4‰ at the base of the Falciferum Ammonite Zone of the by a warming trend in the Early Toarcian, on the basis of oxygen
Penne section. In the ANDRA HTM 102 Borehole (NE France, Paris isotope analyses on bulk rock, or on biogenic carbonates of belemnite
Basin), a prominent negative excursion of δ13C measured on organic rostra and brachiopod shells.
matter occurs in Lower Toarcian sediments deposited after a hiatus;
the base of the carbon excursion parallels an increase in wt.% TOC 4.2. Calcareous nannofossils
(van Breugel et al., 2006). The record of the Tournadous section in the
Causses Basin is therefore very similar to what is reported from two 4.2.1. Preservation
other sedimentary basins belonging to the NW European shelf. These In (palaeo)environmental studies, calcareous nannofossil preser-
similar records make us think that the isotope values measured in the vation has to be carefully checked in order to avoid misinterpretations
Tournadous section are primary, and that they represent the of assemblage compositions. In the Upper Pliensbachian sediments,
expression of the Early Toarcian anoxic event, albeit high TOC values calcareous nannofossils are generally well-preserved at Tournadous
persisted in the Causses Basin until the end of the Early Toarcian. (Fig. 6) and moderately preserved at Saint-Paul-des-Fonts (Fig. 7). In
The oxygen isotope data (bulk carbonates) show average values general etching, although limited, is more commonly observed than
of −4.25‰ in the samples from the Upper Pliensbachian of the overgrowth. As the samples are let settle for 24 h in water
Tournadous section, and −5.5‰ for the Lower Toarcian samples. These oversaturated with respect to carbonates, calcite re-precipitation
very negative values could be seen as the result of diagenetic overprint and overgrowth on calcareous nannofossils may occur, especially
rather than in terms of palaeotemperature. Alternatively to the during the drying of slides. This is the reason why the samples in
diagenetic hypothesis, the very negative values of oxygen isotopes which overgrowth was observed have been re-analysed using simple
in the Causses Basin may be due to a low-salinity superficial water- smear-slide preparation. Because the quantity of water used for smear
body. Oxygen isotope values in the same range as in the Causses Basin slides is very limited, this technique prevents nannofossil overgrowth
are reported from the Quercy Basin (SW France; Emmanuel et al., during sample preparation. The two sets of samples were then
2006) and from the south-west German Basin for the same time compared, and no difference was observed in nannofossil overgrowth.
354
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364
Fig. 6. Absolute and relative abundance of the most commonly recorded nannofossil taxa in the Tournadous section. Ex. = Exaratum Ammonite Subzone. The shaded area corresponds to the T-OAE. The various taxa show peaks in abundance in
different stratigraphic intervals with respect to the T-OAE.
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364
Fig. 7. Absolute and relative abundance of the most commonly recorded nannofossil taxa in the Saint-Paul-des-Fonts section. The shaded area corresponds to the T-OAE. Albeit the proportion of some taxa is different with respect to TND (see
Fig. 6), a similar trend in nannofossil distribution is observed.

355
356
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364
Fig. 8. Only 10 of the 14 studied samples from the Tournadous section yielded benthic foraminifers. From left to right: relative abundance of fragmented foraminifers and relative abundance of specimens bearing evidence of abrasion, dissolution,
micro-boring; relative abundance of Lenticulina E and Lenticulina D morphogroups; relative abundance of genera belonging to the different morphogroups. See the text for further explanation. The shaded area corresponds to the T-OAE.
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 357

The nodular, carbonate-rich beds of the Late Pliensbachian (probably (Schistes Papyracés), where wt.% CaCO3 is also extremely low (b5%), or
diagenetic in origin) contain moderately or badly preserved calcar- in the nodular, diagenetic limestone beds. We do not recognize in the
eous nannofossils; fragmentation of schizospheres is also observed studied sections the “schizospherellid crisis” inferred for the same
in these levels. The clay-rich base of the Schistes carton (Schistes time interval by Erba (2004) and Tremolada et al. (2005). In fact, not
Papyracés) is characterized in the two sections by poorly to only Schizosphaerella but also coccolith abundance is drastically
moderately preserved nannofossils. Preservation improves along reduced during the T-OAE. The rest of the Early Toarcian shows
with the increase in calcium carbonate content within the Serpenti- increasing calcareous nannofossil abundance up to the highest values
num/Falciferum Ammonite Zone (Figs. 6 and 7). As the overall effects (2200E6 specimens per gram of rock) measured in the Saint-Paul-des-
of etching and overgrowth are limited, and because of the presence of Fonts section. In this interval, carbonate values also increase in each
coccoliths prone to dissolution (such as specimens of the Biscutaceae section (on average, CaCO3 is ~45%).
or Calyculus) all along the studied interval, we can reasonably consider
that the assemblage composition of the Tournadous and Saint-Paul- 4.2.4. Assemblage changes through time
des-Fonts samples is rather pristine. In the Tournadous section, calcareous nannofossil assemblages are
largely dominated by coccoliths, while Schizosphaerella spp. accounts
4.2.2. Biohorizons in the studied sections on average for less than 12% of total assemblages (Fig. 2). Relatively
At Tournadous (Fig. 2), Lotharingius sigillatus and Crepidolithus high proportions of this taxon are recorded only in nodular,
cavus (=C. impontus of Bown and Cooper, 1998) are recorded from carbonate-rich beds that bear poor assemblages with rare coccoliths.
the base of the section, 2.62 m below the Pliensbachian/Toarcian In Fig. 6, absolute and relative abundance of the most commonly
boundary as defined by ammonites. Calyculus spp. is recorded recorded coccoliths show the same general trend. The fluctuations in
from 0.38 m, very close to the base of the studied section. We do relative abundance are therefore not related to a closed sum effect,
not consider therefore this datum as a real first occurrence (FO). with the exception of few samples (i.e., at 3.4 m). Some taxa, such as
Lotharingius crucicentralis and Lotharingius aff. L. velatus first occur at Bussonius (mainly B. prinsii with subordinated B. leufuensis), L. barozii,
1.33 m and 1.54 m, respectively, within the Spinatum Ammonite Zone T. patulus and P. liasicus, have relative abundances showing high
(Late Pliensbachian). The FO of Biscutum aff. B. intermedium is at values at the base of the section, in the Spinatum Ammonite Zone.
1.98 m. At Saint-Paul-des-Fonts (Fig. 2), L. sigillatus, L. crucicentralis, Some other taxa (Crucirhabdus and S. novum + S. finchii) are relatively
L. aff. velatus, C. cavus, and B. aff. intermedium are recorded from the abundant from the late Spinatum Zone to the base of the Elegans
base of the section, 2.66 m below the Pliensbachian/Toarcian Subzone. It must be observed, however, that between these two zones
boundary. Lotharingius aff. L. velatus, very similar to L. velatus but there is a gap including at least all the Tenuicostatum Zone. It is
with a thinner ring, has been already reported by Veiga de Oliveira interesting to note that in the Late Pliensbachian under-calcified
et al. (2005) from the latest Pliensbachian of the Lusitanian Basin coccoliths of Bussonius, Lotharingius and Similiscutum are commonly
(Peniche section). recorded. A peak in absolute and relative abundance of Calyculus and
The majority of the stratigraphic interval corresponding to the Orthogonoides is observed at the base of the Elegans Zone. This last
Tenuicostatum Ammonite Zone is lacking in the entire Causses Basin form is an incertae sedis that morphologically resembles to ascidia
(Trümpy, 1983; Guex et al., 2001; Morard et al., 2003). However, in the spicules of the living didemnidae (see Kott, 2005). The species
Saint-Paul-des-Fonts section a horizon is recorded at 2.6 m that is dominating the assemblage in the upper part of the section (Elegans
enriched in ammonite specimens characteristic of the upper part of the Subzone) are Similiscutum spp., L. hauffii, C. crassus and C. cavus.
Semicelatum Subzone (Morard, 2004). At Tournadous, the hiatus In the Saint-Paul-des-Fonts section the trends of calcareous
seems to be more important than at Saint-Paul-des-Fonts because the nannofossil assemblages are similar to those observed at Tournadous.
base of the Schistes carton is dated as Serpentinum/Falciferum The abundance of Schizosphaerella (Fig. 2) is lower than at Tournadous.
Ammonite Zone (Morard, 2004). The FOs of Carinolithus superbus, As the Late Pliensbachian interval at Saint-Paul-des-Fonts is shorter
C. poulnabronei, and Discorhabdus ignotus are recorded in the same in terms of duration than that studied at Tournadous, Bussonius,
sample at Tournadous, whilst at Saint-Paul-des-Fonts D. ignotus is L. barozii, T. patulus, P. liasicus, Crucirhabdus and S. novum +S. finchii are
recorded some 20 cm above Carinolithus. Accordingly, calcareous observed to peak in the same ammonite zone. In the interval where
nannofossil data also suggest a different duration of the hiatus in the calcareous nannofossil absolute abundance is very low, Calyculus and
two sections. Mitrolithus jansae has not been observed at Tournadous, Orthogonoides display a peak (Exaratum to Elegans Ammonite Sub-
and only rare and sporadic specimens have been observed at Saint- zones) whereas, in the rest of the Early Toarcian, assemblages are
Paul-des-Fonts, which was situated in a more distal position within the dominated by Similiscutum spp., L. hauffii, C. crassus and C. cavus. It is
Causses Basin. The last occurrence (LO) of this taxon is recorded interesting to note that Crepidolithus crassus, which is very abundant in
at 6.9 m (Fig. 2). On the basis of calcareous nannofossil record, it is the Elegans Ammonite Subzone in both sections, is the largest coccolith
possible to recognize in the two sections the nannofossil Subzone NJ5b in Lower Toarcian assemblages. Although entire coccospheres have not
C. cavus (=C. impontus), as defined for the NW Europe by Bown and been recorded, some coccolith measurements performed in this work
Cooper (1998). The base of the overlaying calcareous nannofossil Zone allow us to estimate an average volume of 163 µm3 for the specimens
NJ6 C. superbus very probably lays within the sedimentary hiatus. of the Causses Basin, which compared to modern coccolithophore
species is quite huge. The main differences between the two
4.2.3. Calcareous nannofossil absolute abundance sections are related to the percentage of various taxa (see for example
On average, nannofossil abundance (sum of coccoliths and L. barozii; Figs. 6 and 7) that can be more or less abundant in proximal
nannoliths) is higher in the Saint-Paul-des-Fonts section (352E6 spec- (Tournadous) with respect to distal settings (Saint-Paul-des-Fonts).
imens per gram of rock) than at Tournadous (287E6 specimens per
gram of rock; Fig. 2). Coccoliths are dominant over the nannolith 4.3. Benthic foraminifers
Schizosphaerella (graphs with empty dots in Fig. 2) in both sections.
Abundance values are quite constant and relatively low in the Late In the Late Pliensbachian and Early Toarcian interval of the
Pliensbachian of Tournadous, where wt.% CaCO3 is also low (b20%). Tournadous section, the benthic foraminiferal assemblages are largely
Absolute abundances are relatively high in both sections in the latest dominated (98–99%) by hyaline taxa (mainly calcitic Lagenida and
part of the Pliensbachian, Hawskerense Ammonite Subzone, where aragonitic Robertinida). Porcellanaceous and agglutinated taxa are
wt.% CaCO3 varies between 30% and 80%. Very low abundance values very rare (1–2%) in the studied samples (Fig. 8). Foraminifers are
or barren samples are recorded in the basal part of the Toarcian generally moderately to poorly preserved. Most of the tests are filled
358 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

with secondary calcite, recrystallised, and a high percentage of oxygenation at the sea bottom are important factors controlling
specimens are fragmented (from 10 to 20% in the Upper Pliensbachian benthic life. The parameters of bottom-water oxygenation and organic
samples; Fig. 8). In the uppermost part of the Pliensbachian and in the flux (food availability) are closely interrelated, and it is often extremely
Lower Toarcian samples, the original tests are often replaced by pyrite difficult to distinguish their respective influences (Jorissen, 1999). In
(for example, from the sample TND 21 upwards in the Tournadous the samples TND 1, 5 and 9, foraminiferal assemblages are char-
section). Although we cannot exclude a preservation control on the acterized by abundant, highly sculptured Marginulina and by
foraminifer assemblage composition, some patterns of faunal assem- finely ornamented Ichthyolaria. Both taxa can attain a large size
blages seem to correspond to primary environmental signals, such as (250–1000 μm). Large Lenticulina (135–625 μm) are also recorded in
variations in hydrodynamic energy, flux of organic matter, oxygen the same levels (Fig. 8). Morphogroups A, C, and D are relatively
level, etc. abundant in this interval (Fig. 8). Foraminiferal assemblage is
dominated by infauna in the samples TND 1 and 5, whereas infauna
4.3.1. Hydrodynamism vs. dissolution is more or less equivalent to epifauna in TND 9 (Fig. 3). These data may
In the Tournadous section (Fig. 8), the relative abundance curve be indicative of conditions favourable to benthic life development, a
of the fragmented specimens is roughly parallel to that of the robust, quite stable and oxidizing oxidising environment. However, an
biconvex, and involute Lenticulina (morphogroup E). From samples environmental stress was probably induced by a progressive increase
TND 9 to TND 29, the high relative abundance of both fragmented in hydrodynamic energy from sample TND 1 to TND 9, as discussed
specimens and Lenticulina morphogroup E may be interpreted as above. The reduction in the number of genera in the sample TND 9
the result of a selective dissolution. However, high abundance of could be, at least in part, the result of this increase in energy.
Lenticulina morphogroup E and of fragmented specimens do not The sample TND 13 is characterized by a peak in the relative
coincide with the increase in siliceous agglutinated taxa, suggesting abundance of Lenticulina morphogroup D and of Reinholdella, and by a
that foraminiferal assemblages still preserve primary, environmental marked decrease in relative abundance of Ichtyolaria and Marginulina
information. As an alternative explanation, the increase in Lenticulina (Fig. 8). Interestingly, this change in the assemblage composition is
E together with fragmented specimens may be interpreted as the correlated with a little peak of TOC (0.90%) and of HI (140), as shown in
result of important winnowing. The increase in signs of abrasion on Fig. 3. Reinholdella and Lenticulina morphogroup D become dominant
the Lenticulina tests (Fig. 8), especially in the first Toarcian sample in the Early Toarcian, in the interval characterized by levels enriched
(TND 29), supports the hypothesis of an important winnowing. in organic matter (TND 37 and 41), and probably corresponding to
An exception is represented by the sample TND 37, where a dysaerobic conditions at the sea bottom. The assemblage of the sample
peak of Lenticulina E corresponds to an increase in dissolution of the TND 13 likely mirrors an increase in the organic-matter flux (Fig. 8). In
Lenticulina tests, but in this case a decrease of fragmented specimens spite of the increase in the organic-matter flux, sea bottom was (at least
was observed. In this sample, the dissolution has clearly a major temporarily) well oxygenated, as indicated by the highest genus
impact on the assemblage composition. Interestingly, this peak of richness observed in this sample (Fig. 8), and by the comparable
Lenticulina E with signs of dissolution is associated with abundant abundance of infauna and epifauna (Fig. 3). Moreover, almost all the
(20%) Reinholdella, an aragonitic taxon that should be very susceptible foraminiferal morphogroups are present in this sample (Fig. 8). Cool
to dissolution. We may explain this paradox assemblage as the result water temperatures in this interval are probably responsible for a
of “time averaging” (Martin, 1999). Lenticulina and Reinholdella were higher concentration of dissolved oxygen in bottom waters. Oxidising
probably not co-existing, but they likely correspond to a succession of bottom waters likely favoured organic-matter remineralization.
different assemblages, adapted to peculiar environmental conditions. In the upper part of the Spinatum Ammonite Zone, below the
These assemblages are recorded in an artificial mixing, due to the sedimentary hiatus (samples TND 21 and TND 25), the faunal
extremely low accumulation rate and condensation. Robust, involute composition changed as shown by the important increase in relative
and biconvex Lenticulina (morphogroup E) were likely better adapted abundance of Paralingulina (20–25%), associated with high propor-
to increased energy with respect to the more evolute, flat Lenticulina tions of Marginulina (18–25%), and a marked decrease in relative
and Astacolus (morphogroup D; see also Larsen and Drooger, 1977; abundance of Ichthyolaria and Lenticulina morphogroup D. We
Hallock, 1979; Rey et al., 1994). observed in these samples a general reduction in size of benthic
The increased proportion of morphogroup E with respect to foraminifers, and especially of Marginulina (b250 μm). It is noteworthy
morphogroup D (E / E + D × 100) in the samples TND 1 to TND 25 that Ichthyolaria and Paralingulina are roughly anti-correlated
(Fig. 8) seems also be the result of a high environmental energy, throughout the section (Fig. 8).
probably related to the regressive trend of the end-Pliensbachian. In the Umbria–Marche Basin (central Italy; South-West Tethys),
This general trend of increased energy is interrupted by a decrease of Paralingulina has been recorded dominant in the Lower Toarcian
energy in the sample TND 13. After the emersion leading to the black shales, where wt.% TOC attains its highest values (1.4–2.2%;
Pliensbachian/Toarcian hiatus (Morard et al., 2003), the Lower Bartolini et al., 1992). The increase in relative abundance of Paralin-
Toarcian transgression produced shallow-sea environments in gulina at Tournadous may be related to an increase in nutrients and
which storm events occurred, as testified by the levels bearing organic-matter flux, but the sea-bottom conditions still remain
hummocky cross-stratifications. High energy and shallow-water oxidising as indicated by the high genus richness. According to the
conditions are also suggested by the highest proportion of the model of Jorissen et al. (1995), Paralingulina tenera could be
morphogroup E with respect to morphogroup D in the sample TND considered as a deep infaunal taxon, favoured by a high flux of
29 (Fig. 8). In this sample, a small peak (1.4%) of Ophtalmidium and organic matter (eutrophisation), and tolerant to dysoxic conditions. In
the presence of rare Nubecularia (fixed specimens) may indicate the Upper Pliensbachian samples of Tournadous, the increase in
shallow-water conditions, extreme condensation, and the temporary nutrients is associated with a regressive phase and a probable cooling
development of firm grounds. Stratigraphically upwards (TND 37 trend, as indicated by the δ18O curve (Fig. 3). The slight increase in
and 41), the proportion of morphogroup E vs. morphogroup D relative abundance of siliceous agglutinated Glomospirella in the
decreases (Fig. 8), probably mirroring a decrease in environmental sample TND 25, just below the hiatus, may be related with this cooling
energy and a deepening trend. trend or to a slight variation in ambient pH (Nocchi and Bartolini,
1994).
4.3.2. Organic Organic-matter flux and sea-bottom oxygenation The first sample of the Early Toarcian above the hiatus (TND 29) is
Besides the control of hydrodynamic energy and of sea-level characterized by a slight decrease in the genus richness (Fig. 8). The
changes on foraminiferal assemblages, organic-matter flux and faunal composition is characterized by an increase in relative
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 359

abundance of Lenticulina morphogroup E and by the presence of in press). Surface waters in the Causses Basin seem therefore to be cool
porcellanaceous Ophtalmidium. The relative abundance of small-sized and rich in nutrients in the Late Pliensbachian (Fig. 9a). Nutrients were
Paralingulina and Marginulina (b250 μm) is still high with respect to also available at sea bottom, as indicated by relatively high propor-
the samples in the upper part of the Spinatum Ammonite Zone. The tions of Paralingulina. This taxon was in fact part of the deep infauna,
sample TND 29 corresponds to a marked increase in organic matter and was favoured by a high organic flux (Bartolini et al., 1992). The
(4.27%; Fig. 3). In the overlying samples (TND 37 and TND 41), a water column at that time was well oxygenated as testified by low
substantial change in the assemblage composition is observed, organic-matter preservation in sediments, and by the abundant and
namely a dramatic decrease of the genus richness (Fig. 8). Epifauna diverse benthic foraminifer assemblages (Fig. 9).
is dominant over infauna (Fig. 3), foraminifers are significantly small A clear regressive trend is documented in the Late Pliensbachian of
and non-ornamented, the diversity is reduced and the morphogroup the Causses Basin (Fig. 9b), culminating with emersion at the end-
G (mainly Reinholdella) becomes progressively dominant (Fig. 8). The Pliensbachian/base of Toarcian (Guex et al., 2001; Morard et al., 2003).
samples TND 37 and TND 41 are relatively enriched in organic matter Similarly, in the Lusitanian Basin (Portugal) an important regression is
(wt.% TOC is 1.95 and 1.88, respectively), and they are in an interval inferred for the end-Pliensbachian on the basis of facies analysis
characterized by very high TOC values (6–8%). We interpret the (Duarte et al., 2004, 2007). A regressive trend and an increase in
Reinholdella as opportunistic forms, tolerant to hypoxic conditions environmental energy are inferred for the Tournadous section on
(see also Copestake and Johnson, 1981). Reinholdella was probably the basis of foraminiferal assemblage (Fig. 8). Biconvex, quite globular
an opportunistic phytodetritus feeder, surface-dwelling species, Lenticulina morphogroup E are in fact observed that, due to sea-
which could benefit of short periods of bottom-water re-oxygenation. bottom winnowing, are often abraded. The record of low temperatures
In the samples TND 29 and TND 37, the high values of wt.% TOC and in a period of global sea-level regression (Hallam, 2001; Hardenbol
the co-existence of taxa indicating oxidising environments, such as et al., 1998) leading to emersion and hiatuses in various NW Europe
Lenticulina morphogroup E, with low-oxygen tolerant taxa, such as settings (i.e., South Burgundy area, Elmi and Rulleau, 1993; Causses
Paralingulina and Reinholdella, may be the result of “time averaging”, Basin, Guex et al., 2001; Provence area, Floquet et al., 2003; Paris Basin,
as discussed previously. In this interval, the faunal composition pro- van Breugel et al., 2006; sill between the Paris and Aquitanian Basins,
bably indicates alternating oxidising and hypoxic conditions. Alter- Galbrun et al., 1994; Courtinat et al., 2007), is consistent with the
natively dysoxic to anoxic sediments are likely to have favoured hypothesis of ice-caps formation (Guex et al., 2001; Morard et al.,
carbonate dissolution when re-oxygenation occurs (Green et al., 2003). The occurrence of glendonites and dropstones at high
1993). This could explain the presence of robust Lenticulina tests with palaeolatitudes in the Late Pliensbachian (Price, 1999) further
clear trace of dissolution, as the remains of selective dissolution supports this hypothesis.
processes during re-oxygenated intervals.
The general trend of foraminiferal assemblage at Tournadous 5.2. Palaeoenvironmental conditions during the T-OAE
indicates that the environmental conditions deteriorated in terms of
sea-bottom oxygenation passing from the Late Pliensbachian to the A warming trend characterizes the Early Toarcian (McArthur et al.,
Early Toarcian, until a complete disappearance of benthic fauna from 2000; Bailey et al., 2003; Rosales et al., 2004; van de Schootbrugge
4.45 m upwards. Foraminifers are absent also in the finest fraction et al., 2005a,b; Suan et al., 2008a), that it is probably related to an
(50 μm). This level probably corresponds to the development of more increase in atmospheric pCO2 (Hesselbo et al., 2000; Beerling et al.,
permanent sea-bottom anoxia. In the Saint-Paul-des-Fonts section, all 2002; Suan et al., 2008a), although it is not excluded that temperature
the analysed Lower Toarcian samples are barren in benthic forami- fluctuations were modulated by orbital cycles (Suan et al., 2008b).
nifers, probably indicating that in this deeper area of the basin the sea- This warming trend occurred at the same time as a global, 3rd order
bottom anoxia developed earlier than in the shallower Tournadous transgression (Hardenbol et al., 1998; Hallam, 2001), although the
area. precise relationship between warming and sea-level changes has not
been ascertained yet. The transgression is recorded diachronously in
5. Discussion the Causses area, according to the morphology of the basin. At Saint-
Paul-des-Fonts (more distal) a level enriched in ammonite specimens
5.1. Palaeoenvironmental conditions prior to the T-OAE indicating the Semicelatum Ammonite Subzone (corresponding to a
part of the 1st ammonite zone of the Early Toarcian) is recorded
In spite of little differences in the abundance of some calcareous after the hiatus (Fig. 2) (Morard, 2004). Conversely, at Tournadous
nannofossils recorded in the proximal (Tournadous section) with (more proximal) the first ammonite horizons are indicative of the
respect to the distal (Saint-Paul-des-Fonts section) settings of the second ammonite zone of the Toarcian, the Exaratum Ammonite
Causses Basin, some common trend can be traced. Some taxa, such Subzone (Fig. 2). Calcareous nannofossil biostratigraphy further
as P. liasicus, Bussonius spp., and L. barozii, have relative and abso- confirms the diachroneity in the record of the transgression, as it
lute abundance peaks in the Late Pliensbachian of both sections can be seen in Fig. 2 with the different occurrence of C. poulnabronei,
(Figs. 6 and 7). These taxa are usually recorded in higher proportions C. superbus and D. ignotus in the two sections. Although the Upper
along the northern margin of the Tethys than in sections at low Pliensbachian/Lower Toarcian hiatus had a different temporal exten-
palaeolatitudes (Bucefalo Palliani et al., 2002; Mattioli et al., in press). sion within the basin, the T-OAE is recorded in the two sections.
Their latitudinal distribution can be interpreted in terms of sea- In many sections of the western Tethys, the T-OAE is not only
surface temperature, and their abundance in the Late Pliensbachian associated to an enrichment in organic matter in the sediments, but
may therefore indicate the development of temperate water-masses also to a negative carbon isotope excursion occurring in the Exaratum
at relatively low palaeolatitudes. This interpretation is consistent with Ammonite Zone (Hesselbo et al., 2000). At Tournadous (this work) and
the general trend of cooling recorded from the Late Pliensbachian Saint-Paul-des-Fonts (Morard, 2004), the T-OAE corresponds to the
on the basis of oxygen isotopes or Mg/Ca ratio measured on the interval in which very negative δ13C values (~−4‰) and a first increase
carbonates of belemnites (McArthur et al., 2000; Bailey et al., 2003; in TOC are recorded (shaded area in Figs. 3 and 4). TOC values remain
Rosales et al., 2004; van de Schootbrugge et al., 2005a,b) and of oxygen high in the two sections for the rest of the Early Toarcian. However,
isotopes of brachiopod shells (Suan et al., 2008a). This interval is it must be noted that high TOC values are commonly recorded in
also characterized by the occurrence of meso-eutrophic nannofossil sediments overlying the T-OAE, mainly in sections located along the
taxa (namely, S. novum and S. finchii; Bucefalo Palliani and Mattioli, N-Tethyan margin (Baudin, 1989; see also the Dotternhausen section
1995; Mattioli and Pittet, 2004; Tremolada et al., 2005; Mattioli et al., in SW Germany, Röhl et al., 2001, and the HTM-102 Borehole in NE
360 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

Fig. 9. Cartoon showing sea level and oxygenation evolution along a N–S transect in the Causses Basin. (a) During the Late Pliensbachian, the entire water column was well
oxygenated, nutrients were available both at sea bottom and in surface waters. (b) Between the latest Pliensbachian and the earliest Toarcian, a low sea level led to emersion of many
localities of the Causses Basin. (c) The following trend to increasing temperatures might have led to a sea-level rise, a stratification for density and, finally, to the development of
anoxia within the Causses Basin. Benthic fauna is absent in the studied sections, and nannofossil assemblages are dominated by taxa adapted to stressing conditions (see text for
further explanations). (d) With the ongoing sea-level rise and the development of storms, water stratification and anoxia were progressively reduced. Surface water environments
were re-colonised by an abundant and diverse nannoplankton community. Benthic foraminifers developed again at the basin margins, whilst in the central parts of the basin sea
bottom was still anoxic.

France, van Breugel et al., 2006). Within some sedimentary basins, events developed within the Causses basin, as demonstrated by storm
stratification of the water column and sea-bottom anoxia persisted deposits showing HCS. These events probably produced short-lived
long-time after the T-OAE, at least until the end of the Early Toarcian re-oxygenation at the sea bottom. Benthic foraminifers in fact indicate
or even in the Middle Toarcian (Baudin, 1989; Baudin et al., 1990). alternating dysoxic–oxic conditions at the sea bottom in the first
Apparently, anoxic conditions persisted in spite of the fact that storm Toarcian deposits of Tournadous, but environmental conditions
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 361

rapidly deteriorate in terms of sea-bottom oxygenation during the T- These bacteria use the magnetic properties of magnetite or griegite
OAE until a complete disappearance of benthic fauna (Fig. 5). At Saint- under dysoxic and anoxic conditions, respectively, to be oriented with
Paul-des-Fonts, benthic fauna is completely absent in the lower respect to oxidising zones within the water column (Kopp, 2001). In
Toarcian samples. This datum may indicate that sea-bottom anoxia the majority of magnetotactic bacteria, these crystals form chains or
developed earlier in the deeper parts of the Causses Basin and clusters (Bazylinski and Frankel, 2000). The record of such clusters in
eventually migrated towards its margins (Fig. 9). the Serpentinum/Falciferum Zones of the studied sections may testify
Different authors have indicated that the base of the negative for the occurrence of reducing conditions within the water column
isotope excursion roughly coincides with the base of the C. superbus (Fig. 9c and d). In Lower Toarcian black shales from different Tethyan
nannofossil Zone (Bailey et al., 2003; Mattioli and Pittet, 2004; settings, anoxia within the photic zone has been inferred on the basis
Mattioli et al., 2004; Tremolada et al., 2005; Mailliot et al., 2006; van of the presence of biomarkers derived from photosynthesing bacteria
Breugel et al., 2006). The same record is observed in this work. living under sulphide, anoxic conditions (Schouten et al., 2000;
The lowest absolute abundances of calcareous nannofossils are Pancost et al., 2004; Schwark and Frimmel, 2004; van Breugel et al.,
recorded in the Exaratum Ammonite Subzone in both studied 2006). If reducing conditions attained per ascensum the base of
sections. Although nannofossil preservation is moderate or poor in the photic zone in the Causses Basin, deep-dweller coccoliths (like
the samples from the Exaratum Ammonite Subzone, these very low C. crassus; Bour et al., 2007; Mattioli et al., in press) could not develop
abundances are not interpreted here as being related to nannofossil during the T-OAE (Fig. 9c). This is probably the reason why the lowest
dissolution in the water column or within the sediment. In fact, abundance of Crepidolithus spp. is recorded in this interval.
solution-susceptible coccoliths (like Calyculus spp., or the small and
delicate Crucirhabdus spp.) show some abundance peaks in this 5.3. Palaeoenvironmental conditions after the T-OAE
interval. Furthermore, this pattern is not uniquely observed in the
Causses Basin, a similar drastic decrease in abundance is observed in Absolute nannofossil abundances attain very high values in the
various settings of the western Tethys (Mattioli et al., 2004; Erba, interval above the T-OAE (Fig. 9d). This is a common pattern in western
2004; Tremolada et al., 2005; Mattioli et al., in press). Tethys (Bucefalo Palliani et al., 2002; Mattioli and Pittet, 2004; Mattioli
Finally, similarly to the record of many other Tethyan settings et al., 2004; Tremolada et al., 2005; Mattioli et al., in press). This pattern
(Mattioli et al., in press), during the T-OAE a peak in Calyculus spp. is testifies for a recovery of environmental conditions within the photic
observed in this work concomitant with very low absolute abundance of zone, and an improvement of trophic conditions. Stratification for
nannofossils. In both sections, high relative and absolute abundances density was probably removed thanks to the storm action and,
of Calyculus spp. are recorded along with a peak in Orthogonoides although nutrients were available all the time as the Causses Basin
hamiltoniae (shaded area in Figs. 6 and 7). Calyculus spp. is considered as was surrounded by emerged lands (representing source areas), the re-
inhabiting the uppermost surface waters, and being adapted to extreme occurring mixing within the water column favoured a better re-cycling
surface water conditions, in terms of low salinity and water column of nutritive elements. Nutrients were also available at the sea bottom,
stratification (Mattioli et al., in press) Low salinity in the surface waters as indicated by the high proportions in this time interval of the benthic
of the Causses Basin is consistent with the anomalously negative oxygen foraminifer Reinholdella (Tournadous section). This taxon is commonly
isotope values measured in the Tournadous samples. interpreted as an opportunist, sediment surface-dweller, which likely
Although previous works (Mattioli et al., 2004; Erba, 2004; proliferated during short-lived periods of sea-bottom re-oxygenation.
Tremolada et al., 2005) interpret the drastic decrease in abundance After the T-OAE, the recolonisation by nannoplankton of the deep
of calcareous nannofossils during the T-OAE as a marine bio- photic zone occurred due to the removal of anoxia, according to the
calcification crisis directly induced by high atmospheric pCO2, high relative and absolute abundance of the deep-dweller Crepido-
the results of this work rather validate the hypothesis of Mattioli lithus crassus (Figs. 6 and 7). Anoxia persisted however in the deepest
et al. (in press). The latter authors consider the crisis of calcareous parts of the Causses Basin, because benthic foraminifers did not
nannoplankton and the dominance of tolerant taxa during the T-OAE recover in the Saint-Paul-des-Fonts section (Fig. 9d).
(mainly Calyculus spp. and Orthogonoides hamiltoniae), as an indirect There is a fairly good parallelism between the nannofossil absolute
consequence of high atmospheric CO2. In fact, climatic effects of abundance increase, the dominance in the assemblage of Crepidolithus
increased CO2 are an enhancement of the hydrological cycle (Cohen spp., and the highest values of TOC and HI (compare Figs. 6 and 7 with
et al., 2004; Hesselbo et al., 2007), increase of fresh-water inputs from Figs. 4 and 5). The nature of the organic matter is essentially marine
continents to the seas (Röhl et al., 2001), and an important lowering of in this interval, and primary productivity seems to be elevated. High
salinity in surface waters of epicontinental basins surrounding the primary productivity is also inferred on the basis of important rela-
Tethys (McArthur et al., 2000; Röhl et al., 2001; van de Schootbrugge tive and absolute abundances of meso-eutrophic coccolith taxa
et al., 2005a; Emmanuel et al., 2006). Besides the environmental stress (mainly Similiscutum finchii and S. novum; Bucefalo Palliani et al.,
induced by a low-saline surface waterbody, the peaks in Calyculus spp. 2002; Mattioli and Pittet, 2004; Tremolada et al., 2005; Mattioli et al.,
and Orthogonoides spp. in a period of very low proliferation of other in press). Although an interpretation of the carbon isotope trend
nannoplankton taxa can be also related to the presence of reduced during the Early Toarcian is behind the scope of this work, a high
nitrates within the photic zone. Indeed, a positive shift of δ15N in black primary productivity starting from the Elegans Ammonite Subzone
shales of the T-OAE (Jenkyns et al., 2001) is interpreted in terms of would be consistent with the positive excursion of δ13C. A high
bacterial denitrification within the water column and diffusion of NH+4. primary production leads to a decrease in the 12C reservoir of surface
In these conditions, the majority of coccolithophorids were probably waters because of photosynthesis, resulting in a positive excursion of
overwhelmed by other phytoplankton groups (Bucefalo Palliani et al., δ13C of marine carbonates (Anderson et Arthur, 1983). We postulate
2002; van de Schootbrugge et al., 2005a,b). that nannoplankton was a major contributor to marine productivity at
In the interval enriched in organic matter, pyrite framboids are the end of the Early Toarcian, and that Crepidolithus crassus, the largest
commonly recorded. The cubic or tetrahedral crystals associated to the coccolith in Lower Toarcian assemblages, could have significantly
framboids probably correspond to magnetite or griegite. If these contributed to biomass production.
magnetite or griegite crystals are primary, they could have been
produced by biomineralisation of magnetotactic bacteria (e.g., 6. Summary and conclusions
Aquaspirillum magnetotacticum, see Frankel and Blakemore, 1984),
living in marine environments at the transition zone between In geological times characterized by important organic-matter
oxidising and reducing conditions (Bazylinski and Frankel, 2000). accumulation, like during OAEs, it is difficult to assess the contribution
362 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

of an enhanced primary production and, respectively, bottom-water Acknowledgements


reducing conditions enabling the preservation of organic matter
in sediments. In the present paper, an integrated study (calcareous We wish to warmly thank Alessandra Negri for inviting us to
nannofossils, benthic foraminifers, and geochemistry) of two sections submit a contribution to the special issue “Organic carbon rich
located in a restricted basin characterized by a limited water-depth sediment through the Phanerozoic: Processes, Progresses and Per-
(the Causses Basin; Fig. 9) enabled us to reconstruct the fluctuations in spectives”. The constructive remarks of Michaël Hermoso and an
the nutrient concentration of the whole water column (i.e., in surface anonymous reviewer greatly improved the quality of this manuscript.
waters and at the sea bottom), the photic zone structure, and the F. Savignac is gratefully acknowledged for her assistance in Rock-Eval
deep-water oxygenation during the T-OAE. The geological record of analyses. Martine Fordant, Liana Goedde, Frédérique Bellec and
this event is characterized in the Causses Basin, as well as in various Gueorgui Ratzov are thanked for their precious help for foraminiferal
other localities, by sediments depleted in CaCO3 but enriched in and geochemical analyses. We enjoyed the discussions with Serge
organic matter, by a negative excursion of carbon isotopes, the virtual Elmi about stratigraphy all along this work. His advice and remarks
absence of benthic fauna, and a drastic decrease in abundance of were invaluable to us. Publication No. UMR 5125-08.xxx.
calcareous nannofossils.
Nutrients were likely available all the time in the studied interval
within the Causses Basin. Indeed this basin was surrounded by References
emerged lands that represented the source areas for nutrients. In spite
of the nutrient availability, nannoplankton production greatly varied Anderson, T.F., Arthur, M.A., 1983. Stable isotopes of oxygen and carbon and their
in the considered time interval. Calcareous nannofossil assemblages application to sedimentologic and paleoenvironmental problems. In: Arthur, M.A.,
Anderson, T.F. (Eds.), Stable Isotopes in Sedimentary Geology. SEPM Short Course,
are characterized by relatively high fertility index (Biscutaceae) in the vol. 10, pp. 1.1–1.151.
Late Pliensbachian and late Early Toarcian, namely prior to and after Arthaud, F., Matte, P., 1975. Les décrochements tardi-hercyniens du Sud-Ouest de
the T-OAE. Similarly, small but significant peaks of opportunistic l'Europe. Géométrie et essai de reconstitution des conditions de la déformation.
Tectonophysics 25, 139–171.
foraminifers, adapted to high trophic conditions, are recorded in the Bailey, T.R., Rosenthal, Y., Mc Arthur, J.M., van de Schootbrugge, B., Thirlwall, M.F., 2003.
two periods. However, benthic fauna indicates that during the Late Paleoceanographic changes of the Late Pliensbachian–Early Toarcian interval: a
Pliensbachian, the sea bottom was well oxygenated. Such conditions possible link to the genesis of an Oceanic Anoxic Event. Earth and Planetary Science
Letter 212, 307–320.
likely prevented organic-matter preservation in the sediment. Cool Bartolini, A., Nocchi, M., Baldanza, A., Parisi, G., 1992. Benthic life during the Early
water temperatures in the Late Pliensbachian are probably responsible Toarcian Anoxic Event in the Southwestern Tethyan Umbria–Marche basin, Central
for a higher concentration of dissolved oxygen in bottom waters, and Italy. Studies in Benthic Foraminifers, Benthos '90, Sendai. Tokai University Press,
pp. 323–338.
subsequent organic-matter remineralization. An important hiatus
Bassoulet, J.P., Elmi, S., Poisson, A., Cecca, F., Bellion, Y., Guiraud, R., Baudin, F., 1993. Mid
spanning the topmost Pliensbachian and the basal Toarcian is Toarcian. In: Dercourt, J., Ricou, L.E., Vrielynck, B. (Eds.), Atlas Tethys paleoenvironmental
documented for the Causses Basin. This hiatus is associated with maps. BEICIP — FRANLAB, Rueil-Malmaison, pp. 63–84.
low sea level and emersions in many basins surrounding the western Baudin, F., 1989. Caractérisation géochimique et sédimentologique de la matière
organique du Toarcien téthysien (Méditerranée, Moyen-Orient), significations
Tethys and might have had a glacio-eustatic origin, according to paléogeographiques. Thèse, Université Pierre et Marie Curie (Paris VI), pp. 1–246.
Guex et al. (2001) and Morard et al. (2003). The following global Baudin, F., Herbin, J.P., Vandenbroucke, M.,1990. Mapping and geochemical characterization
warming occurring in the Early Toarcian was coupled to an enhanced of the Toarcian organic matter in the Mediterranean Tethys and Middle East. Advances
in Organic Geochemistry 16, 677–687.
hydrological cycle and riverine input, promoting the formation of low- Bazylinski, D., Frankel, R., 2000. Magnetic iron oxide and iron sulfide minerals within
salinity water-bodies and, ultimately, a severe stratification for density microorganisms. In: Baeuerlein, E. (Ed.), Biomineralization: from biology to biotechnology
and widespread anoxia in all the epicontinental basins of the NW and medical application. Wiley-VCH, Weinheim, Germany, pp. 25–46.
Beerling, D.J., Lomas, M.R., Gröcke, D.R., 2002. On the nature of methane gas-hydrate
Europe, including the Causses Basin. These conditions were favourable dissociation during the Toarcian and Aptian oceanic anoxic events. American
to organic-matter accumulation and preservation in sediments. In the Journal of Science 302, 28–49.
Causses area, anoxia developed at the sea bottom first in the central Bernhard, J.M., 1986. Characteristic assemblages and morphologies of benthic
foraminifera from anoxis, organic-rich deposits: Jurassic through Holocene. Journal
part of the basin, then the oxygen minimum zone expanded laterally of Foraminiferal Research 16, 207–215.
toward the shallower areas, and vertically towards the photic zone. Bour, I., Mattioli, E., Pittet, B., 2007. Nannofacies analysis as a tool to reconstruct
Anoxia development represented the most critical condition for paleoenvironmental changes during the Early Toarcian anoxic event. Palaeogeography,
Palaeoclimatology, Palaeoecology 249, 58–79.
benthic organisms that temporarily disappeared during the T-OAE.
Bown, P.R., Cooper, M.K.E., 1998. Jurassic. In: Bown, P.R. (Ed.), Calcareous Nannofossil
At the same time, surface water environments underwent very Biostratigraphy. . British Micropalaeontological Society Publication Series. Kluver
stressing conditions testified by a severe drop in nannofossil abun- Academic Publishers, Cambridge, pp. 33–85.
dance, the temporary disappearance of taxa thriving in the lower Bucefalo Palliani, R., Mattioli, E., 1995. Ecology of dinoflagellate cysts and calcareous
nannofossils from bituminous facies of the Early Toarcian, central Italy. Proceedings
photic zone (like C. crassus), and peaks of taxa able to thrive in low- III EPA Workshop: Black Shales Models, Dotternhausen, pp. 60–62.
salinity waters enriched in reduced nitrates (like Calyculus). Bucefalo Palliani, R., Mattioli, E., Riding, J.B., 2002. The response of marine
At the end of the Early Toarcian, stratification and anoxia were phytoplancton phytoplankton and sedimentary organic matter to the early
Toarcian (Lower Jurassic) oceanic anoxic event in northern England. Marine
removed from surface waters because of the development of storms, and Micropaleontology 46, 223–245.
deep-dweller nannoplankton could thrive again. In the basin margins, Buzas, M.A., Culver, S.J., Jorissen, F.J., 1993. A statistical evaluation of the microhabitats of
anoxic conditions at the sea bottom were probably alternating with living (stained) infaunal benthic foraminifera. Marine Micropaleontology 20, 311–320.
Cecca, F., Macchioni, F., 2004. The two Early Toarcian (Early Jurassic) extinction events in
short-lived re-oxygenation, as indicated by the occurrence of abundant, Ammonoids. Lethaia 37, 35–56.
opportunist epifaunal foraminifers. However, in the deeper and confined Cohen, A.S., Coe, A.L., Harding, S.M., Schwark, L., 2004. Osmium isotope evidence for the
parts of the Causses Basin anoxia persisted as demonstrated by the regulation of atmospheric CO2 by continental weathering. Geology 32, 157–160.
Copestake, P., Johnson, B., 1981. The Hettangian to Toarcian. In: Jenkyns, D.G., Murray, J.W.
absence of benthic foraminifers in the more distal Saint-Paul-des-Fonts (Eds.), Stratigraphical Atlas of Fossil Foraminifera. British Micropaleontological Society,
section. The highest absolute abundance of nannofossils was recorded Ellis Horwood Limited, pp. 81–105.
after the T-OAE, along with high proportions of the coccolith taxa of Corliss, B.H., 1985. Microhabitats of benthic foraminifera within deep-sea sediments.
Nature 314, 435–438.
largest size. This datum seems to indicate that the contribution of
Corliss, B.H., 1991. Morphology and microhabitat preferences of benthic foraminifera
calcareous phytoplankton to the organic-matter fraction, that is mainly from the northwest Atlantic Ocean. Marine Micropaleontology 17, 195–236.
marine in this interval, was significant. The results of this work indicate Corliss, B.H., Chen, C., 1988. Morphotype patterns of Norwegian Sea deep-sea benthic
that, although this interpretation has been recently challenged foraminifera and ecological implications. Geology 16, 716–719.
Courtinat, B., Pittet, B., Mattioli, E., Rio, M., 2007. Organic walled dinoflagellate cyst and
(McArthur, 2007), the T-OAE was a global event that has been recorded acritarch assemblages compared to nannofossil abundance and palaeoenviron-
even in small and partly enclosed basins (like the Causses Basin). mental context in the Toarcian stratotype. Géobios 40, 785–800.
S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364 363

Duarte, L.V., Wright, V.P., Fernández-Lopez, S., Elmi, S., Krauter, M., Azerêdo, A., Kopp, 2001. Biomineralization in magnetotactic bacteria. College paper (University of
Henriques, M.H., Rodrigues, R., Perilli, N., 2004. Early Jurassic carbonate evolution in Chicago) 28–11-2001.
the Lusitanian Basin (Portugal): facies, sequence stratigraphy and cyclicity. In: Kott, P., 2005. New and little-known species of Didemnidae (Ascidiacea, Tunicata) from
Duarte, L.V., Henriques, M.H. (Eds.), Carboniferous and Jurassic carbonate platforms Australia (part 3). Journal of Natural History 39, 2409–2479.
of Iberia, Fieldtrip guidebook 23rd IAS Meeting of Sedimentology, Coimbra 2004, Larsen, A.R., Drooger, C.W., 1977. Relative thickness of the tests in the Amphistegina
vol. 1, pp. 47–71. species of the Gulf of Elat. Utrecht Micropaleontological Bulletin 30, 225–240.
Duarte, L.V., Oliveira, L.C., Rodrigues, R., 2007. Carbon isotopes as a sequence Little, C.T.S., Benton, M.J., 1995. Early Jurassic mass extinction: a global long-term event.
stratigraphic tool: examples from the Lower and Middle Toarcian marly limestones Geology 23, 495–498.
of Portugal. Boletín Geológico y Minero 118, 3–18. Loeblich, A.R., Tappan, J.H. 1988. Foraminiferal Genera and their classification. Van
Elmi, S., Rulleau, L., 1993. Le Jurassique du Beaujolais Meridional, bordure orientale du Nostrand Reinhold Company New York, 2nd volume, text volume 970 p.; plates
Massif Central, France. Geobios 15, 139–155. volume, 212 p. and 847 plates.
Emmanuel, L., Renard, M., Cubaynes, R., de Rafélis, M., Hermoso, M., LeCallonec, L., Le Mailliot, S., Mattioli, E., Guex, J., Pittet, B., 2006. The Early Toarcian anoxia, a synchronous
Solleuz, A., Rey, J., 2006. The “Schistes Carton” of Quercy (Tarn, France): a event in the Western Tethys? An approach by quantitative biochronology (Unitary
lithological signature of a Methane hydrate dissociation event in the early Toarcian. Associations), applied on calcareous nannofossils. Palaeogeography, Palaeoclimatol-
Implications for correlations between Boreal and Tethyan realms. Bulletin Societé ogy, Palaeoecology 240, 562–586.
géologique de France 5, 239–249. Martin, R.E., 1999. Taphonomy and temporal resolution of foraminiferal assemblages. In:
Erba, E., 2004. Calcareous nannofossils and Mesozoic oceanic anoxic events. Marine Sen Gupta, B.K. (Ed.), Modern Foraminifera. Kluwer Academic Publishers, pp. 281–298.
Micropaleontology 52, 85–106. Mattei, J., Elmi, S., Mouterde, R., Tintant, H., Gabilly, J., 1967. Le Domérien dans quelques
Espitalié, J., 1993. Rock Eval pyrolysis. In: Bordenave, M.L. (Ed.), Applied petroleum régions du Centre et du Sud de la France. Colloque du Jurassique, Luxembourg.
geochemistry, Technip, Paris, pp. 237–261. Mémoire du Bureau de Recherche Géologique et Minière 75, 567–580.
Espitalié, J., Deroo, G., Marquis, F., 1985. La pyrolyse Rock Eval et ses applications Mattioli, E., Pittet, B., 2004. Spatial and temporal distribution of calcareous nannofossils
(deuxième partie). Revue de l'Institut Français du Pétrole 40, 755–784. along a proximal–distal transect in the Lower Jurassic of the Umbria–Marche Basin
Farrimond, P., Englinton, G., Brassell, S.C., Jenkyns, H., 1989. Toarcian anoxic event in (central Italy). Palaeogeography, Palaeoclimatology, Palaeoecology 205, 295–316.
Europe: an organic geochemical study. Marine and Petrology Geology 6, 136–147. Mattioli, E., Pittet, B., Bucefalo Palliani, R., Röhl, H.J., Schmid-Röhl, A., Morettini, E., 2004.
Floquet, M., Cecca, F., Mestre, M., Macchioni, F., Guiomar, M., Baudin, F., Durlet, C., Alméras, Y., Phytoplankton evidence for the timing and correlation of palaeoceanographical
2003. Mortalité en masse ou fossilisation exceptionnelle? Le cas des gisements d'âge changes during the early Toarcian oceanic anoxic event (Early Jurassic). Journal of
toarcien inférieur et moyen de la région de Digne-Les-Bains (Sud-Est de la France). the Geological Society London 161, 685–693.
Bulletin de la Societé géologique de France 174, 159–176. Mattioli, E., Pittet, B., Suan, G., Mailliot, S., in press. Calcareous nannoplankton changes
Frankel, R., Blakemore, R., 1984. Precipitation of Fe3O4 in magnetotactic bacteria. across the early Toarcian oceanic anoxic event in the bwestern Tethys. Paleoceano-
Philosophical Transactions of the Royal Society of London B 304, 567–574. graphy. 23. doi:10.1029/2007PA001435.
Galbrun, B., Baudin, F., Bassoulet, J.-P., Depeche, F., Emmanuel, L., Lachkar, G., Renard, M., McArthur, J.M., 2007. Comment on “Carbon-isotope record of the Early Jurassic
Riveline, J., Gabilly, J., Hantzpegue, P., Manivit, H., Rouget, C.,1994. Stratigraphie intégrée (Toarcian) Oceanic Anoxic Event from fossil wood and marine carbonate
du Toarcien stratotypique (coupes de Thouars et Airvault, Deux-Sèvres, France). (Lusitanian Basin, Portugal)” by Hesselbo S., Jenkyns H.C., Duarte L.V. and Oliveira
Geobios 17, 575–595. L.C.V. Earth and Planetary Science Letters 259, 634–639.
Geisen, M., Bollmann, J., Herrle, J.O., Mutterlose, J., Young, J.R., 1999. Calibration of the McArthur, J.M., Donovan, D.T., Thirvall, M.F., Fouke, B.W., Mattey, D., 2000. Strontium
random settling technique for calculation of absolute abundances of calcareous isotope profile of the early Toarcian (Jurassic) oceanic anoxic event, the duration of
nannoplankton. Micropaleontology 45, 437–442. ammonite biozones, and belemnite paleotemperatures. Earth and Planetary
Green, M.A., Aller, R.C., Aller, J.Y., 1993. Carbonate dissolution and temporal abundances of Science Letters 179, 269–285.
Foraminifera in Long Island sediments. Limnology and Oceanography 38, 331–345. Mc Crea, J.M., 1950. On the isotopic chemistry of carbonates and a paleotemperature
Guex, J., 1972. Répartition biostratigraphique des ammonites du Toarcien moyen de la scale. Journal of Chemical Physics 18, 849–857.
bordure sud des Causses (France) et révision des ammonites décrites et figurées par Monestier, J., 1921. Le Toarcien supérieur de la région Sud-Est de l'Aveyron. Bulletin de la
Monestier (1931). Eclogae geologicae Helvetiae 65, 611–645. Société géologique de France 4, 280–312.
Guex, J., Morard, A., Bartolini, A., Morettini, E., 2001. Découverte d'une importante Morard, A., 2004. Les événements du passage Domérien-Toarcien entre Téthys
lacune stratigraphique à la limite Domérien-Toarcien: implications paléo- occidentale et Europe du Nord-Ouest. Thèse, Université Lausanne, Géologie -
océanographiques. Bulletin de la Société Vaudoise des Sciences Naturelles 345, Paléontologie, pp. 1–158.
277–284. Morard, A., Guex, J., Bartolini, A., Morettini, E., de Wever, P., 2003. A new scenario for the
Hallam, A., 2001. A review of the broad pattern of Jurassic sea-level changes and their Domerian–Toarcian transition. Bulletin de la Société géologiquede France 174, 351–356.
possible causes in the light of current knowledge. Palaeogeography, Palaeoclima- Murray, J.W., 2001. The niche of benthic foraminifera, critical thresholds and proxies.
tology, Palaeoecology 167, 23–37. Marine Micropaleontology 41, 1–7.
Hallock, P., 1979. Trends in test shape with depth in large, symbiont-bearing Nguyen, T.M.P., Speijer, R.P., 2007. Experimental dissolution studies on late Mesozoic and
foraminifera. Journal of Foraminiferal Research 9, 61–69. Cenozoic foraminiferal assemblages. Integrated Studies of Taxonomy, Ecology and
Hardenbol, J., Thierry, J., Farley, M., Jacquin, T., De Graciansky, P.-C., Vail, P.R.,1998. Mesozoic Geochemistry, the Micropalaeontological Society's Foraminifera and Nannofossil
and Cenozoic sequence chronostratigraphic framework of European basins. In: De Groups Joint Spring Meeting 2007, Angers (France), p. 100. abstract book.
Graciansky, P.-C., Hardenbol, J., Jacquin, T., Vail, P. (Eds.), Mesozoic and Cenozoic Nocchi, M., Bartolini, A., 1994. Investigation of Lagenina and Glomospirella assemblages
Sequence Stratigraphy of European Basins. SEPM Special Publication, vol. 60, pp. 3–13. in the Umbria–Marche basin in Central Italy during the Late Domerian–early
Harries, P.J., Little, C.T.S., 1999. The early Toarcian (Early Jurassic) and the Cenomanian– Toarcian. 3rd International Symposium on Jurassic Stratigraphy, Poitiers, France.
Turonian (Late Cretaceous) mass extinctions: similarities and contrasts. Palaeogeogra- Geobios, vol. 17, pp. 689–699.
phy, Palaeoclimatology, Palaeoecology 154, 39–66. Pálfy, J., Smith, P.L., 2000. Synchrony between Early Jurassic extinction, oceanic event,
Hesselbo, S.P., Gröcke, D.R., Jenkyns, H.C., Bjerrum, C.J., Farrimond, P., Morgans Bell, H.S., and the Karoo–Ferrar flood basalt volcanism. Geology 28, 747–750.
Green, O.R., 2000. Massive dissociation of gas hydrate during a Jurassic oceanic Pancost, R.D., Crawford, N., Magness, S., Turner, A., Jenkyns, H.C., Maxwell, J.R., 2004.
event. Nature 406, 392–395. Further evidence for the development of photic-zone euxinic conditions during
Hesselbo, S.P., Jenkyns, H.C., Duarte, L.V., Oliveira, L.C.V., 2007. Carbon-isotope record of Mesozoic oceanic anoxic events. Journal of the Geological Society London 161, 2004,
the Early Jurassic (Toarcian) Oceanic Anoxic Event from fossil wood and marine 1–12.
carbonate (Lusitanian Basin, Portugal). Earth Planetary Science Letters 253, Patterson, W.P., Walter, L.M., 1994. Depletion of 13C in seawater ∑CO2 on modern
455–470. doi:10.1016/j.epsl.2007.05.002. carbonate platforms: significance for the carbon isotopic record of carbonates.
Jenkyns, H.C., 1988. The early Toarcian (Jurassic) Oceanic Anoxic Event: stratigraphic, Geology 22, 885–888.
sedimentary, and geochemical evidence. American Journal of Science 288, 101–151. Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed
Jenkyns, H.C., Clayton, C.J.,1986. Black shales and carbon isotopes in pelagic sediments from pyrolysis. American Association of Petroleum Geologists Bulletin 70, 318–329.
the Tethyan Lower Jurassic. Sedimentology 33, 87–106. Price, G.D., 1999. The evidence and implication of polar ice during the Mesozoic. Earth-
Jenkyns, H.C., Clayton, C.J., 1997. Lower Jurassic epicontinental carbonates and mudstones Science Reviews 48, 183–210.
from England and Wales: chemostratigraphic signals and the Early Toarcian anoxic Raiswell, R., 1988. Chemical model for the origin of minor limestone-shale cycles by
event. Sedimentology 44, 687–706. anaerobic methane oxidation. Geology 16, 641–644.
Jenkyns, H.J., Gröcke, D.R., Hesselbo, S.P., 2001. Nitrogen isotope evidence for water mass Rey, J., Bonnet, L., Cubaynes, R., Qajon, A., Ruget, C., 1994. Sequence stratigraphy and
denitrification during the early Toarcian (Jurassic) oceanic anoxic event. Paleoceano- biological signals: statistical studies of benthic foraminifera from Liassic series.
graphy 16, 593–603. Palaeogeography, Palaeclimatology, Palaeoecology 111, 149–171.
Jenkyns, H.C., Jones, C.E., Gröcke, D.R., Hesselbo, S.P., Parkinson, D.N., 2002. Chemostrati- Röhl, H.J., Schmid-Röhl, A., Oschmann, W., Frimmel, A., Schwark, L., 2001. The Posidonia
graphy of the Jurassic System: applications, limitations and implications for Shale (Lower Toarcian) of SW Germany: an oxygen-depleted ecosystem controlled
palaeoceanography. Journal of the Geological Society London 159, 351–378. by sea level and palaeoclimate. Palaeogeography, Palaeclimatology, Palaeoecology
Jorissen, F.J., De Stigter, H.C., Widmark, J.G.V., 1995. A conceptual model explaining benthic 169, 273–299.
foraminiferal microhabitats. Marine Micropaleontology 26, 3–15. Rosales, I., Quesada, S., Robles, S., 2004. Paleotemperature variations of Early Jurassic
Jorissen, F.J., 1999. Benthic foraminiferal microhabitats below the sediment–water seawater recorded in geochemical trends of belemnites from the Basque–Cantabrian
interface. In: Sen Gupta, B.K. (Ed.), Modern Foraminifera. Kluwer Academic Publishers, basin, northern Spain. Palaeogeography, Palaeoclimatology, Palaeoecology 203,
pp. 161–179. 253–275.
Kälin, O., Bernoulli, D., 1984. Schizosphaerella Deflandre and Dangeard in Jurassic deep Roth, P.H., 1984. Preservation of calcareous nannofossils and fine-grained carbonate
water carbonate sediments, Mazagan continental margin (Hole 547B) and particles in mid-Cretaceous sediments from the southern Angola Basin, site 530.
Mesozoic Tethys. Initial Reports of the Deep Sea Drilling Project 79, 411–435. Initial Reports of the Deep Sea Drilling Project 75, 651–655.
364 S. Mailliot et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 346–364

Sass, E., Bein, A., Almogi-Labin, A., 1991. Oxygen-isotope composition of diagenetic Trümpy, D.M., 1983. Le Lias Moyen et Supérieur des Grands Causses et de la région de
calcite in organic-rich rocks: evidence for 18O depletion in marine anaerobic pore Rodez: contributions stratigraphiques, sédimentologiques et géochimiques à la
water. Geology 19, 839–842. connaissance d'un bassin à sédimentation marneuse. Cahiers de l'Université,
Scholle, P.A., Arthur, M.A., 1980. Carbon isotope fluctuations in cretaceous pelagic Université de Pau et des Pays de l'Adour 19, 1–363.
limestones: potential stratigraphic and petroleum exploration tool. American van Breugel, Y., Schouten, S., Baas, M., Mattioli, E., Sinninghe Damsté, J.S., 2006.
Association of Petroleum Geologists Bulletin 64, 67–87. Isorenieratane record in black shales from the Paris Basin, France: constraints on
Schouten, S., Kaam-Peters, M.E., Rijpstra, I., Schoell, M., Sinnighe Damste, J.S., 2000. recycling of respired CO2 as a mechanism for negative carbon isotope shifts during
Effects of an oceanic anoxic event on the stable carbon isotopic composition of early the Toarcian oceanic anoxic event. Paleoceanography 21, PA4220. doi:10.1029/
Toarcian carbon. American Journal of Science 300, 1–22. 2006PA001305.
Schwark, L., Frimmel, A., 2004. Chemostratigraphy of the posidonia Black Shale, SW van de Schootbrugge, B., McArthur, J.M., Bailey, T.R., Rosenthal, Y., Wright, J.D., Miller, K.G.,
Germany II. Assessment of extent and persistence of photoc-zone anoxia using aryl 2005a. Toarcian oceanic anoxic event: an assessment of global causes using belemnite
isoprenoid distribution. Chemical Geology 206, 231–248. C isotope records. Paleoceanography 20, PA3008. doi:10.1029/2004PA001102.
Suan, G., Mattioli, E., Pittet, B., Mailliot, S., Lécuyer, C., 2008a. Evidence for major van de Schootbrugge, B., Bailey, T.R., Rosenthal, Y., Katz, M.E., Wright, J.D., Miller, K.G.,
environmental perturbation prior to and during the Toarcian (Early Jurassic) Feist-Burkhardt, S., Falkowski, P.G., 2005b. Early Jurassic climate change and the
Oceanic Anoxic Event from the Lusitanian Basin, Portugal. Paleoceanography 23, radiation of organic walled phytoplankton in the Tethys Ocean. Paleobiology 31,
PA1202. doi:10.1029/2007PA001459, 2008. 73–97.
Suan, G., Pittet, B., Bour, I., Mattioli, E., Duarte, L.V., Mailliot, S., 2008b. Duration of the Veiga de Oliveira, L.C., Perilli, N., Duarte, L.V., 2005. Calcareous nannofossil assemblages
Early Toarcian carbon isotope excursion deduced from spectral analysis: conse- around the Pliensbachian/Toarcian Stage Boundary in the reference section of Peniche
quence for its possible causes. Earth and Planetary Science Letters 267, 666–679. (Portugal). In: Elmi, S., Duarte, L., Mouterde, R., Rocha, R.B., Soares, A.F. (Eds.), The
doi:10.1016/j.epsl.2007.12.017. Peniche Section (Portugal) Candidate to the Toarcian Global Stratotype Section and
Swart, P.K., Burns, S.J., Leder, J.J., 1991. Fractionation of the stable isotopes of oxygen and Point (GSSP), Toarcian working group field trip meeting, Caparica, pp. 39–45.
carbon dioxide during the reaction of calcite with phosphoric acid as a function of Wignall, P.B., 2001. Large igneous provinces and mass extinctions. Earth-Science
temperature and technique. Chemical Geology 86, 89–96. Reviews 53, 1–33.
Tissot, B.P., Welte, D.H., 1984. Petroleum Formation and Occurrence, 2nd edition. Wignall, P.B., Newton, R.J., Little, C.T.S., 2005. The timing of paleoenvironmental change
Springer-Verlag, Berlin. 699 pp. and cause-and-effect relationships during the Early Jurassic mass extinction in
Tremolada, F., van de Schootbrugge, B., Erba, E., 2005. Early Jurassic schizosphaerellid Europe. American Journal of Science 305, 1014–1032.
crisis in Cantabria, Spain: implications for calcification rates and phytoplankton Wignall, P.B., Newton, R.J., Reeves, E., Hallam, A., Mattioli, E., Sha Jingeng, N., Crowley, S.,
evolution across the Toarcian oceanic anoxic event. Paleoceanography 20, PA2011. 2006. An eastern Tethyan (Tibetan) record of the Early Jurassic (Toarcian) mass
doi:10.1029/2004PA01120. extinction event. Geobiology 4, 179–190.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 365–367

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Abstract of “The role of sea-level change and marine anoxia in the


Frasnian–Famennian (Late Devonian) mass extinction”☆
David P.G. Bond ⁎, Paul B. Wignall
School of Earth and Environment, University of Leeds, Leeds, LS2 9JT, United Kingdom

Introduction Johnson et al. (1985, p. 578) considered the sea-level rise component
of cycle IId to be:
The Frasnian–Famennian mass extinction (F–F, Late Devonian) is one
“the greatest of Devonian transgressions… (because it) coincides with
of the “big 5” faunal crises of the Phanerozoic with taxa being lost from a
the West Falls Group of New York and encompasses the Kellwasser
broad range of marine habitats (Hallam and Wignall, 1997). Johnson et
Limestone of Germany and the Matagne Shale of Belgium… (and)
al. (1985) proposed one of the first explicit links between marine anoxia,
comprises a pair of widely recognised transgressions”.
transgression and mass extinction for the F–F mass extinction, a cause-
and-effect nexus which has been accepted by many. Extinction losses of The two transgressions were separated by “a small-scale drop in
groups such as the ostracods, conodonts, and tentaculitoids are sea level” (Johnson et al., 1985, p. 584) and were followed by a major
contemporaneous with the widespread deposition of anoxic facies, regression in the Middle triangularis Zone. The first of the transgres-
most notably the Upper Kellwasser Horizon of Germany, and many sions occurred within the Lower gigas Zone and is contemporaneous
workers have attributed the extinction to this phenomenon (e.g. with the development of the Lower Kellwasser Horizon in Germany.
Joachimski and Buggisch, 1993; Becker and House, 1994; Levman and Unfortunately, Johnson et al. (1985) provided conflicting ages for the
von Bitter, 2002; Bond et al., 2004). Others (e.g. Copper, 1986; Sandberg second transgression and thus sowed the seeds of confusion in much
et al., 1988, 2002) prefer sea-level fall and cooling as an extinction of the subsequent literature. In their figure 12 the second transgres-
mechanism. The relationship between sea-level, the Upper Kellwasser sion was shown as beginning at the base of the Lower triangularis
anoxic event and the contemporaneous mass extinction is a subject of Zone, but they state in their text that this transgression correlates with
conflicting opinions (e.g. Hallam and Wignall, 1999 versus Sandberg et the Upper Kellwasser Horizon. This began in the Uppermost gigas
al., 2002). Here, we examine the validity of the F–F boundary portion of Zone as correctly shown in their time-rock chart (Johnson et al., 1985,
the classic Johnson et al. (1985) curve using facies analyses of sections figure 2). Johnson et al. (1985, p. 581) clearly stated that “the Frasnian–
studied by the authors and recently-published data from the literature in early Famennian transgressive history supports an interpretation that
order to critically assess the role of sea-level change during the mass a succession of three rapid deepening events within and above IId, not
extinction and its relationship with contemporary redox changes. regression, caused many of the Frasnian extinctions”. We therefore
assume that their figure 12 was poorly drafted and that the second
The Johnson et al. (1985) eustatic sea-level curve transgression of T–R Cycle IId coincides with the development of the
Upper Kellwasser Horizon in the Uppermost gigas Zone, and thus with
The Johnson et al. (1985) eustatic sea-level curve identified two the F–F mass extinction (Fig. 1).
major “depophases” (termed I and II) within the Devonian, each Significant changes in the Late Devonian conodont zonation scheme
consisting of 6 transgressive–regressive (T–R) cycles labelled a to f. have occurred since 1985. The Lower to Upper gigas interval is now
The Pragian to Frasnian was a time of rising sea-level, with the late approximated by the Early to Late rhenana Zones, whilst the Uppermost
Frasnian being a period of second-order highstand, before sea-level gigas Zone has become the linguiformis Zone (Ziegler and Sandberg,
began to fall in the Famennian. T–R Cycle IId is of relevance here, 1990). The F–F boundary has also been redefined (Sandberg et al., 1988).
because this cycle straddles the F–F mass extinction interval (Fig. 1). In 1985 it was placed at the Lower/Middle triangularis zonal boundary
but it is now placed at the base of the Lower (now more correctly called
Early) triangularis Zone. Thus, the second major transgression of the
Johnson et al. (1985) T–R cycle IId now begins within the linguiformis
Zone and the major regression at the top of the cycle is well within the
Famennian rather than at the old F–F boundary (Fig. 1).

☆ The full version of this paper has been published in: Palaeogeography, Worldwide evidence for sea-level change
Palaeoclimatology, Palaeoecology 263, 2008, 107–118.
DOI of original article: 10.1016/j.palaeo.2008.02.015.
⁎ Corresponding author. Fax: +44 113 3435259. New facies analysis of sections in the USA (Nevada, Utah, New York
E-mail address: d.bond@see.leeds.ac.uk (D.P.G. Bond). State) and Europe (France, Germany, Poland), and comparison with

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.12.012
366 D.P.G. Bond, P.B. Wignall / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 365–367

Fig. 1. The eustatic sea-level curve of Johnson et al. (1985), on the left as reproduced in their figure 12, and on the right as described in their text. Note Lower (L), Upper (U), and
Uppermost (Um) gigas Zones are now replaced by Early and Late rhenana and linguiformis Zones respectively. Lower (L), Middle (M) and Upper (U) triangularis Zones are now more
correctly termed Early, Middle, and Late triangularis Zones.

sections known from the literature in Canada, Australia and China Since the publication of the original version of this manuscript, the
reveal several high-frequency relative sea-level changes in the late latest Exxon Palaeozoic sea-level curve has been published (Haq and
Frasnian to earliest Famennian extinction interval. A clear signal of Schutter, 2008). Although by necessity this curve is produced at a
major transgression is seen within the Early rhenana Zone in the Great somewhat lower resolution than our study, it clearly indicates two
Basin of the western United States (e.g. drowning of the carbonate transgressions in the late Frasnian, separated by a small-scale drop in
platform, recorded as a transition from fossiliferous limestones to sea-level, and terminated by major regression in the earliest
laminated micrites in the Devils Gate section, and from fossiliferous Famennian. Thus, the Haq and Schutter (2008) sea-level curve
limestones to laminated shales in the Coyote Knolls section). This correlates well with the history described above.
transgression is also recorded in the eastern United States by the
development of the Pipe Creek Shale above the Nunda Sandstone, and Anoxia and mass extinction
in western Europe at the onset of Lower Kellwasser anoxia. This is the
base of transgressive–regressive Cycle IId of the Johnson et al. (1985) The close association of transgression, the development of anoxic
eustatic curve. facies, and the Late Devonian mass extinction has lead many authors
The transgression at the base of Cycle IId was curtailed by to attribute a cause-and-effect relationship (e.g. Buggisch, 1972;
regression and sequence boundary generation within the early lin- House, 1985; Casier, 1987; Geldsetzer et al., 1987; Goodfellow et al.,
guiformis Zone, recorded by hardground and karstification surfaces in 1989; Walliser et al., 1989; Buggisch, 1991; Becker, 1993; Joachimski
sections from Canada to Australia. This major eustatic fall probably and Buggisch, 1993; Becker and House, 1994; Joachimski et al., 2001,
terminated platform carbonate deposition over wide areas, especially 2002; Levman and von Bitter, 2002; Chen and Tucker, 2003; Bond
in western North America. et al., 2004; Tribovillard et al., 2004; Bond and Wignall, 2005; Riquier
The second of the pair of transgressions in Cycle IId begins in the et al., 2005; Bond, 2006; Pujol et al., 2006). Sea level does not change
later linguiformis Zone, and is recorded by the development of in isolation within the earth-surface system and it is likely that the
organic-rich shale or bituminous limestone facies in sections as major eustatic changes associated with F–F mass extinction indicate
widespread as the western and eastern United States (e.g. Whiterock destabilisation of the climate and C cycle (e.g. Copper, 1986; Buggisch,
Canyon and Beaver Meadow Creek respectively), Canada (the Moose 1991; Joachimski and Buggisch, 1993; Becker and House, 1994; Algeo
River Basin of Ontario (Levman and von Bitter, 2002)), France (e.g. et al., 1995; Algeo and Scheckler, 1998; Streel et al., 2000; Joachimski
Coumiac and La Serre), Poland (Kowala Quarry), and Germany et al., 2002; Goddéris and Joachimski, 2004; Averbuch et al., 2005;
(Steinbruch Benner). In South China, this transgression is recorded Chen et al., 2005; Riquier et al., 2005).
by the infilling of a prominent palaeokarst surface with dark grey The transgression–anoxia–extinction scenario invoked here and by
limestones of late linguiformis Zone age (Chen and Tucker, 2003, those authors cited above appears to be a pattern which was repeated
2004). In Australia, two late Frasnian transgressions are manifest in several times during the Devonian. Brett and Baird (1995) recognised
the Canning Basin as the development of upper marginal-slope facies six Ecological–Evolutionary subunits in the Early Devonian to Frasnian
in reef-margin settings at a time of backstepping stratal geometry, interval, at least five of which were apparently terminated by
and, using carbon isotope stratigraphy, these have recently been dated widespread anoxic highstands. Thus, there were probably several
as contemporaneous with the Kellwasser transgressions in Europe lesser extinctions during the Devonian, and the Frasnian–Famennian
(Stephens and Sumner, 2003). event was merely a more intense manifestation of this scenario.
This linguiformis Zone transgression is significant because it is
associated with the expansion of anoxic deposition, known as the Upper Acknowledgements
Kellwasser Event. Johnson et al.'s (1985) original transgression–anoxia–
extinction link is thus supported, although some extinction losses of The authors wish to thank Steve Kershaw and an anonymous
platform carbonate biota during the preceeding regression cannot be reviewer for their helpful comments on the original version of this
ruled out. Claims for a latest Frasnian regression (e.g. Sandberg et al., manuscript. We also would like to thank those who assisted our
2002) are not supported, and probably reflect poor biostratigraphic fieldwork, which was part of the first author's NERC-funded PhD project:
dating of the early linguiformis Zone sequence boundary. Dieter Korn, Jared Morrow, Jeff Over, Greg Racki, and Michal Zaton.
D.P.G. Bond, P.B. Wignall / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 365–367 367

References Goodfellow, W.D., Geldsetzer, H.H.J., McLaren, D.J., Orchard, M.J., Klapper, G., 1989.
Geochemical and isotopic anomalies associated with the Frasnian–Famennian
extinction. Historical Biology 2, 51–72.
Algeo, T.J., Scheckler, S.E., 1998. Terrestrial–marine teleconnections in the Devonian;
Hallam, A., Wignall, P.B., 1997. Mass Extinctions and their Aftermath. Oxford University
links between the evolution of land plants, weathering processes, and marine
Press, New York.
anoxic events. Philosophical Transactions of the Royal Society of London, Series B
Hallam, A., Wignall, P.B., 1999. Mass extinctions and sea-level changes. Earth Science
353, 113–130.
Reviews 48, 217–250.
Algeo, T.J., Berner, R.A., Maynard, J.B., Scheckler, S.E., 1995. Late Devonian oceanic anoxic
Haq, B.U., Schutter, S.R., 2008. A chronology of Paleozoic sea-level changes. Science 322,
events and biotic crises: ‘rooted’ in the evolution of vascular land plants? GSA Today
64–68.
5, 63–66.
House, M.R., 1985. Correlation of mid-Palaeozoic ammonoid evolutionary events with
Averbuch, O., Tribovillard, N., Devleeschouwer, X., Riquier, L., Mistiaen, B., van Vliet-
global sedimentary perturbations. Nature 313, 17–22.
Lanoe, B., 2005. Mountain building-enhanced continental weathering and organic
Joachimski, M.M., Buggisch, W., 1993. Anoxic events in the late Frasnian — causes of the
carbon burial as major causes for climatic cooling at the Frasnian–Famennian
Frasnian–Famennian faunal crisis? Geology 21, 675–678.
boundary (c. 376 Ma)? Terra Nova 17, 25–34.
Joachimski, M.M., Ostertag-Henning, C., Pancost, R.D., Strauss, H., Freeman, K.H., Littke,
Becker, R.T., 1993. Stratigraphische Gliederung und Ammonoideen-Faunen im Nehde-
R., Sinninghe-Damsté, J.S., Racki, G., 2001. Water column anoxia, enhanced
nium (Oberdevon II) von Europa und Nord-Afrika. Courier Forschung-Institut
productivity and concomitant changes in d13C and d34S across the Frasnian–
Senckenberg 155, 1–405.
Famennian boundary (Kowala — Holy Cross Mountains/Poland). Chemical Geology
Becker, R.T., House, M.R., 1994. Kellwasser events and goniatite successions in the
175, 109–131.
Devonian of the Montagne Noire with comments on possible causations. Courier
Joachimski, M.M., Pancost, R.D., Freeman, K.H., Ostertag-Henning, C., Buggisch, W.,
Forschung-Institut Senckenberg 16, 45–77.
2002. Carbon isotope geochemistry of the Frasnian–Famennian transition.
Bond, D., 2006. The fate of the homoctenids (Tentaculitoidea) during the Frasnian–
Palaeogeography, Palaeoclimatology, Palaeoecology 181, 91–109.
Famennian mass extinction (Late Devonian). Geobiology 4, 167–177.
Johnson, J.G., Klapper, G., Sandberg, C.A., 1985. Devonian eustatic fluctuations in
Bond, D., Wignall, P.B., 2005. Evidence for Late Devonian (Kellwasser) anoxic events in
Euramerica. Geological Society of America Bulletin 96, 567–587.
the Great Basin, western United States. In: Over, D.J., Morrow, J.R., Wignall, P.B.
Levman, B.G., von Bitter, P.H., 2002. The Frasnian–Famennian (mid-Late Devonian)
(Eds.), Understanding Late Devonian and Permian–Triassic biotic and climatic
boundary in the type section of the Long Rapids Formation, James Bay Lowlands,
events: towards an integrated approach. . Developments in Palaeontology and
northern Ontario, Canada. Canadian Journal of Earth Sciences 39, 1795–1818.
Stratigraphy, vol. 20. Elsevier, Amsterdam, pp. 225–262.
Pujol, F., Berner, Z., Stüben, D., 2006. Palaeoenvironmental changes at the Frasnian/
Bond, D., Wignall, P.B., Racki, G., 2004. Extent and duration of marine anoxia during the
Famennian boundary in key European sections: chemostratigraphic constraints.
Frasnian–Famennian (Late Devonian) mass extinction in Poland, Germany, Austria
Palaeogeography, Palaeoclimatology, Palaeoecology 240, 120–145.
and France. Geological Magazine 141, 173–193.
Riquier, L., Tribovillard, N., Averbuch, O., Joachimski, M.M., Racki, G., Devleeschouwer,
Brett, C.E., Baird, G.C., 1995. Coordinated stasis and evolutionary ecology of Silurian to
X., El Albani, A., Riboulleau, A., 2005. Productivity and bottom water redox
Middle Devonian faunas in the Appalachian Basin. In: Erwin, D.H., Anstey, R.L.
conditions at the Frasnian–Famennian boundary on both sides of the Eovariscan
(Eds.), New approaches to speciation in the fossil record. Columbia University Press,
Belt: constraints from trace-element geochemistry. In: Over, D.J., Morrow, J.R.,
New York, pp. 285–315.
Wignall, P.B. (Eds.), Understanding Late Devonian and Permian–Triassic Biotic and
Buggisch, W., 1972. Zur Geologie und Geochemie der Kellwasserkalke und ihrer
Climatic Events: Towards an Integrated Approach. . Developments in Palaeontology
begleitenden Sedimente (Unteres Oberdevon). Geology and geochemistry of the
and Stratigraphy, vol. 20. Elsevier, Amsterdam, pp. 199–224.
lowermost upper Devonian Kellwasser Limestone and its associated sediments.
Sandberg, C.A., Ziegler, W., Dreesen, R., Butler, J.L., 1988. Part 3: Late Frasnian mass
Abhandlungen des Hessischen Landesamtes fuer Bodenforschung 62, 1–67.
extinction: conodont event stratigraphy, global changes, and possible causes.
Buggisch, W., 1991. The global Frasnian–Famennian 'Kellwasser Event'. Geologische
Courier Forschung-Institut Senckenberg 102, 263–307.
Rundschau 80, 49–72.
Sandberg, C.A., Morrow, J.R., Ziegler, W., 2002. Late Devonian sea-level changes,
Casier, J.-G., 1987. Etude biostratigraphique et paléoécologique des ostracodes du récif
catastrophic events, and mass extinctions. Geological Society of America Special
de marbre rouge du Hautmont à Vodelée (partie supérieure du Frasnien, Bassin de
Paper 356, 473–487.
Dinant, Belgique). Revue de Paléobiologie 6, 193–204.
Stephens, N.P., Sumner, D.Y., 2003. Late Devonian carbon isotope stratigraphy and sea
Chen, D., Tucker, M.E., 2003. The Frasnian–Famennian mass extinction: insights from
level fluctuations, Canning Basin, Western Australia. Palaeogeography, Palaeocli-
high-resolution sequence stratigraphy and cyclostratigraphy in South China.
matology, Palaeoecology 191, 203–219.
Palaeogeography, Palaeoclimatology, Palaeoecology 193, 87–111.
Streel, M., Caputo, M.V., Loboziak, S., Melo, J.H.G., 2000. Late Frasnian–Famennian
Chen, D., Tucker, M.E., 2004. Palaeokarst and its implication for the extinction event at
climates based on palynomorph analyses and the question of the Late Devonian
the Frasnian–Famennian boundary (Guilin, south China). Journal of the Geological
glaciations. Earth Science Reviews 52, 121–173.
Society of London 161, 895–898.
Tribovillard, N., Averbuch, O., Devleeschouwer, X., Racki, G., Riboulleau, A., 2004. Deep-
Chen, D., Qing, H., Li, R., 2005. The Late Devonian Frasnian–Famennian (F/F) biotic crisis:
water anoxia at the Frasnian–Famennian boundary (La Serre, France): a tectoni-
insights from δ13Ccarb, δ13Corg, and 87Sr/86Sr isotope systematics. Earth and
cally-induced Late Devonian oceanic anoxic event? Terra Nova 16, 288–295.
Planetary Science Letters 235, 151–166.
Walliser, O.H., Groos-Uffenorde, H., Schindler, E., Ziegler, W., 1989. On the Upper
Copper, P., 1986. Frasnian/Famennian extinction and cold-water oceans. Geology 14,
Kellwasser Horizon (boundary Frasnian/Famennian). Courier Forschung-Institut
835–839.
Senckenberg 110, 247–255.
Geldsetzer, H.H.J., Goodfellow, W.D., McLaren, D.J., Orchard, M.J., 1987. Sulfur-isotope
Ziegler, W., Sandberg, C.A., 1990. The Late Devonian standard conodont zonation.
anomaly associated with the Frasnian–Famennian extinction, Medicine Lake,
Courier Forschung-Institut Senckenberg 121, 1–115.
Alberta, Canada. Geology 15, 393–396.
Goddéris, Y., Joachimski, M.M., 2004. Global change in the Late Devonian: modelling the
Frasnian–Famennian short-term carbon isotope excursions. Palaeogeography,
Palaeoclimatology, Palaeoecology 202, 309–329.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Black shale deposition in an Upper Ordovician–Silurian permanently stratified,


peri-glacial basin, southern Jordan
Howard A. Armstrong a,⁎, Geoffrey D. Abbott b, Brian R. Turner a, Issa M. Makhlouf c,
Aminu Bayawa Muhammad b, Nikolai Pedentchouk d, Henning Peters e
a
Palaeozoic Environments Group, Department of Earth Sciences, Durham University, Science Laboratories, South Road, Durham DH1 3LE, UK
b
School of Civil Engineering and Geosciences, Drummond Building, University of Newcastle upon Tyne, Newcastle upon Tyne NE1 7RU, UK
c
Department of Earth and Environmental Sciences, Hashemite University, Zarqa, Jordan
d
Petroleum Reservoir Group (PRG), Department of Geology and Geophysics, University of Calgary, 2500 University Drive NW, Calgary, Alberta, Canada T2N 1N4
e
Research Center for Ocean Margins (RCOM), University of Bremen, Post Box 330 440, 28334 Bremen, Germany

A R T I C L E I N F O A B S T R A C T

Article history: The Lower Palaeozoic (Upper Ordovician–Silurian) succession of North Africa contains one of the world's
Received 1 October 2007 most prolific black shale source rocks, yet the origin of these rocks remains contentious. The black shale of
Received in revised form 8 May 2008 the Batra Formation in Jordan was deposited at high palaeolatitude during rapid Hirnantian to early Silurian
Accepted 15 May 2008
deglaciation. Here we report geological and organic geochemical results that provide evidence for an increase
in photic zone primary productivity during ice melting. The decay of this organic matter through oxidative
Keywords:
Black shales
respiration resulted in euxinia, which enhanced the potential for organic matter preservation. The occurrence
Silurian of isorenieratane in all samples indicates euxinia extended from the photic zone to the sediment water
Peri-glacial interface. The stratified basins and fjords of east Antarctica provide a likely modern analogue.
Jordan © 2008 Elsevier B.V. All rights reserved.

1. Introduction evidence for productivity changes, anoxia/euxinia and the increased


burial of organic matter. Bulk δ13C and total organic carbon (%TOC) are
The deposition of marine black shale and the enhanced storage of taken to reflect productivity changes. The presence of isorenieratane
organic carbon (OC) in the geological record indicate fundamental (XXIII; see Appendix A for structure), a biomarker of green sulphur
changes in the functioning of biogeochemical cycles and their bacteria, is indicative of photic zone euxinia. Likely modern analogues
feedbacks during extreme climate modes and transitions (Beckmann are the seasonally isolated basins of east Antarctica. A similar
et al., 2005a; Page et al., 2007). Our understanding of the response of interdisciplinary approach is necessary to elucidate the nature of
the marine environment during these climate states can only be these deposits elsewhere on the Gondwana margin.
gained from a study of deep time analogues.
The Upper Ordovician–Lower Silurian succession of North Africa 2. Stratigraphical and geological context
and Arabia contains thick (∼ 20 m), organic carbon (OC)-rich (up to
The Lower Palaeozoic succession in southern Jordan includes some
15% total organic carbon (TOC)) black shale, widely known as the “hot
750–800 m of well exposed Ordovician siliciclastic sediments deposited
shales,” which are the source of ∼30% of the world's oil (Lüning et al.,
on the margins of the North African (Gondwana) in terrestrial to subtidal
2000, 2006). The origin of these deposits remains contentious
marginal marine and shelf environments (Amireh et al., 2001; Makhlouf,
(Armstrong et al., 2005, 2006). The lower hot shale overlies glacial
1995; Powell et al., 1994; Fig. 1). During the Late Ordovician Jordan was
and glacio-marine sediments deposited during the Hirnantian glacia-
located in a high latitude, east Gondwana setting, 60° S of the equator
tion (∼445 Ma) and have been linked to either, nutrient enrichment of
(Cocks and Torsvik, 2002), less than 100 km from the margins of a ter-
shallow marine environments during coastal upwelling (Lüning et al.,
restrial ice sheet in northwest Saudi Arabia. This ice sheet was char-
2000) or, freshening by deglacial meltwater (Armstrong et al., 2005).
acterised by two major phases of ice advance and retreat (Vaslet, 1990)
Here we report data from Jordan that confirms the lower hot shale
both marked by erosional unconformities (Vaslet, 1990; Figs. 1 and 2). The
was deposited in a stratified, ice margin basin during Hirnantian to
first major glacial ice incised into permafrost-hardened and glacially
early Silurian deglaciation (Armstrong et al., 2005). We relate deglacial
loaded, Tubeiliyat shoreface and nearshore shelf deposits, preferentially
sea level rise (at Milankovitch timescales) and melt water flux to
excavating NW–SE trending major fault-controlled depressions, cutting a
steep-sided U-shaped valley (Turner et al., 2005). This ice advance
⁎ Corresponding author. correlates with the first glacial advance in northwest Saudi Arabia (Vaslet,
E-mail address: h.a.armstrong@durham.ac.uk (H.A. Armstrong). 1990; Miller and Mansour, 2007; Fig. 3), and was followed by deglaciation,

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.05.005
H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377 369

Fig. 1. Lithostratigraphy and chronostratigraphy for the Ordovician and Silurian of Jordan and Saudi Arabia, showing generalised depositional environments for outcrops in the
Southern Desert region of Jordan (redrawn from Turner et al., 2005). Subdivision of the Ammar Formation into the Lower and Upper Ammar is based on Abed et al. (1993).

Fig. 2. Generalised section of the glacial and deglacial succession in the Southern Desert region of Jordan and northwest Saudi Arabia showing the stratigraphy and sediment fill of the
glacially incised palaeovalley systems. Section A is located 0.5 km southwest of Jebel Umier (29° 34′ N, 35° 53′ E) and Section B is from Jebel Ammar (29° 34′ N, 35° 52′ E). Section C is
from northwest Saudi Arabia and is based on Vaslet (1990); reproduced with permission from Turner et al. (2005).
370 H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377

the base of the Batra Formation in well BG-14) contain the graptolite
Neodiplograptus apographon typical of the middle and upper subzones of
the ascensus-acuminatus biozone (sensu Storch, 1990) and indicate an
earliest Silurian age. Graptolites collected at 42.82 m (4.02 m above base
of formation) and 41.57 m (5.27 m above the base of the formation)
contain Normalograptus parvulus. N. parvulus ranges through the pers-
culptus to acuminatus biozones of Hirnantian to Rhuddanian (Llandovery)
age (Zalasiewicz and Tunnicliff,1994, Fig. 3). These age assignments allow
the correlation of the transgressive fill of the upper palaeovalley in Jordan
with the Hirnantian to early Llandovery (Rhuddanian) global eustatic sea
level rise (Cocks and Rickards, 1988; Loydell, 1998).
Armstrong et al. (2005) concluded the base of the black shale is
coincident with the maximum flooding of the first post-glacial highstand
(cf Lüning et al., 2000). The fill of the upper palaeovalley was considered
as an “expanding puddle” as originally defined by Wignall (1991).

3. Materials and methods

Here we use carbon isotopic, biomarker and Rock-Eval analyses on


Fig. 3. Palaeogeographical reconstruction of eastern Gondwana, during the late Ordovician, black shale from the lower Batra Formation (Jordan) to establish water
showing the ice sheet (shaded) and the location of Jordan and Saudi Arabia (redrawn from
column redox conditions. Core samples were obtained from the
Sutcliffe et al., 2000).
immature (average Tmax value of 419 °C), OC-rich lower 18 m section of
a rise in relative sea level and transgressive filling of the palaeovalley. The the Batra Formation in the type area of Wadi Batn el Ghul (well BG14;
latter is recorded by a thin, reworked bottom lag of glaciofluvial Table 1) from the Southern Desert region of Jordan (29°30′50.4″ N
sandstones, overlain by thick, transgressive, shoreface sandstones. Late 35°57′41″ E; Armstrong et al., 2005).
transgressive filling of the palaeovalley was interrupted by a second and
possibly a third subsidiary glacial advance producing a glacially polished 3.1. Total organic carbon (TOC) and Rock-Eval pyrolysis
and grooved surface with intersecting glacial striations, indicating ice flow
from the west and northwest (Turner et al., 2005). The TOC contents of the dried samples were measured using a
The fourth glacial advance produced a regionally extensive low- calibrated LECO CS-244 elemental analyser. Each sample was analysed in
stand tunnel valley beneath the ice sheet (Turner et al., 2002). This was duplicate and standard material was analysed after every 10 analytical
subsequently preserved as a palaeovalley incised into the lower samples to ensure that the analyser maintained its calibration. Rock-Eval
palaeovalley-fill deposits or, where this is missing, into the top of the pyrolysis was carried out using a Delsi Oil Show Analyser. Each sample
Tubeiliyat Formation. This ice advance correlates with the second was pyrolysed in duplicate so that the mean values of the amounts of
major ice advance in Saudi Arabia (Vaslet, 1990), where the Sarah hydrocarbons detected under the S1 and S2 peaks as well as the
Formation similarly records a complex record of ice advance and temperature Tmax, corresponding to the temperature at which the
retreat (Miller and Mansour, 2007). The pattern of “major” ice advances maximum of the S2 hydrocarbon generation occurs during pyrolysis,
interspersed by smaller-scale events has been interpreted as reflecting could be measured. Standard (5.51% TOC; 0.27 mg HC/g of rock S1;
eccentricity and obliquity moderated ice volume changes (Sutcliffe 13.59 mg HC/g of rock S2; and 430 °C Tmax) and then blank samples were
et al., 2000; Armstrong, 2007). Following melting of the fourth ice pyrolysed under these same conditions to ensure that the measure-
sheet the upper palaeovalley filled with transgressive glaciofluvial ments of the unknown quantities were as precise as possible. The
sandstones, marine shoreface sandstones (Turner et al., 2005) and average standard deviations with respect to S1, S2 and Tmax were
finally black shale of the Batra Formation (Armstrong et al., 2005). 0.05 mg HC/g of rock, 0.43 mg HC/g of rock and 2.3 °C respectively.
The Batra Formation is ∼40–120 m thick, and in the Southern
Desert region of Jordan it conformably overlies the Ammar Formation 3.2. Bulk stable carbon isotope analysis
or disconformably overlies the Tubeiliyat Formation (Fig. 1). The lower
13 12
part of the formation is found in many shallow exploration wells, and C/ C ratios (δ13C) were measured on bulk sediments after removal
was restricted to fault-bounded graben structures (Powell et al., 1994; of the inorganic carbonates with dilute HCl using automated online
Turner et al., 2005). The formation is variably graptolitic and contains combustion followed by conventional isotope ratio-mass spectrometry
sparse thin-shelled bivalves and rare trilobites (Masri, 1988). in a VG TripleTrap and Optima dual-inlet mass spectrometer, with δ13C
In the type area the lower part of the Batra Formation comprises values calculated to the Vienna Peedee belemnite (VPDB) scale using a
17.44 m of monotonous, OC-rich black shale, which in thin section within-run laboratory standard (cellulose, Sigma Chemical prod. no. C-
comprise laminated, black siltstone to dark grey homogeneous 6413) calibrated against NBS-19 and NBS-22. Replicate analysis of well-
claystone couplets. The parallel laminated siltstones are OC-rich, and mixed samples indicated a precision of ±0.1‰ (1 S.D.).
grade upwards into mudstone, with irregular patches of siltstone
(? starved ripples) and homogeneous claystones. The couplets were 3.3. Gas chromatography (GC)/gas chromatography–mass spectrometry
interpreted as distal turbidites by Armstrong et al. (2005). The absence of (GC-MS)
bioturbation indicates anoxic/euxinic bottom water during deposition
(e.g. Droser and Bottjer, 1986). The laminites, characteristic of the whole The powdered black shales were Soxhlet extracted with dichlor-
formation, contain pyrite framboids and marcasite concretions through- omethane/methanol (93:7; v/v) for 48 h. An aliquot of each total extract
out, confirming a euxinic depositional environment (Wignall, 1994). was separated by thin layer chromatography (TLC, Kieselgel 60G, 0.5 mm
Andrews (1991) reviewed the biostratigraphy of this formation from thickness) using light petroleum ether (boiling point is from 40 to 60 °C)
surface exposures and exploration wells and concluded that the entire into aliphatic hydrocarbon, aromatic hydrocarbon and polar fractions.
formation ranged in age from Ashgill (persculptus Biozone) to the mid- The aliphatic and aromatic hydrocarbon fractions were analysed using a
Wenlock. Graptolites from the lower part of the core have been identified Hewlett-Packard HP5890 gas chromatograph (GC) equipped with a
(see also Loydell, 2007). Samples from a depth of 46.62 m (close to flame ionisation detector and a fused silica capillary column
H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377 371

Table 1 Steranes and hopanes were identified using published mass spectra
Late Ordovician black shales from a Batra Formation (southern Jordan) borehole (BG14) and relative retention times (e.g. Peters et al., 2005). Isorenieratane
giving sample # and position in the stratigraphical column
(XXIII) was identified by its mass spectrum and by GC-MS co-injection
Sample Height above base of Batra Formation (m) Depth in core (m) experiments on an HP-1 stationary phase (authentic standard of
8402-77 16.76 30.08 isorenieratane (XXIII) was supplied courtesy of S. Schouten, Nether-
3402-19 15.28 31.56 lands Institute for Sea Research, Den Burg, Netherlands). The GC-MS
3402-18 15.14 31.70
results from the co-injection experiments are presented in Fig. 5.
3402-17 14.84 32.00
3402-16 14.34 32.50
3402-15 12.94 33.90 4. Results
3402-14 11.74 35.10
3402-13 10.94 35.90 The presence of parallel laminations accompanied by an absence
8402-75 9.84 37.00
8402-73 9.52 37.32
of bioturbation throughout the section indicates euxinic bottom
8402-74 9.27 37.57 waters during deposition. All samples typically contain acritarchs
8402-72 8.77 38.07 and graptolites (Keegan et al., 1990) indicating a primarily marine
8402-68 7.77 39.07 phytoplanktonic and zooplanktonic source of Type II kerogen. The
8402-67 7.52 39.32
percentage of total organic carbon (% TOC) increases as a function of
8402-66 7.27 39.57
8402-64 6.77 40.07 height above the base of the Batra Formation (Fig. 6A). This section
8402-63 6.52 40.32 is OC-rich with a stepwise increase in %TOC from ∼ 1% to ∼ 3% and
8402-62 6.27 40.57 from ∼ 3% to ∼ 9% at 6.27 m and 12.94 m respectively above the base
8402-61 6.02 40.82 of the section (Armstrong et al., 2005). Figs. 6B and 7 show that the
8402-60 5.77 41.07
Rock-Eval hydrogen index (HI) values of the samples have a range of
8402-59 5.52 41.32
8402-53 4.02 42.82 156 to 402 mg HC/g TOC (mean value = 283 mg HC/g TOC). The δ13C
8402-52 3.77 43.07 of bulk organic matter show a range of − 30.8 to − 29.6‰ with
fluctuations of up to 0.4‰ and a positive shift of 1.4‰ up the section
with the largest increase (∼ 1‰) apparent in the two uppermost
(30 m × 0.25 mm i.d.) coated with either HP-1 or HP-5 stationary phase samples (Fig. 6D).
(film thickness of 0.25 µm). The carrier gas was hydrogen, and the oven The regular sterane carbon number distributions are such that C29 N C27
temperature was held at 50 °C for 2 min and then heated at a rate of 4 °C/ C27 NN C28, where the average value of the C28/C29 steranes ratio is 0.27,
min to 300 °C, at which it was held for 20 min (Fig. 4). which agrees with previous observations that generally this particular
GC-MS was performed using a Hewlett-Packard HP5890II GC ratio is less than about 0.35 for samples older than Silurian (Grantham and
coupled with a Hewlett-Packard 5972 mass spectrometer (ionising Wakefield, 1988). Both the sterane and 17α-hopane distributions indicate
voltage of 70 eV and with the source temperature at 160 °C). The GC thermally immature organic matter and do not vary significantly
was fitted with a fused silica capillary column (30 m × 0.25 mm i.d.) throughout the section. Maxima in the regular steranes/17α-hopanes
coated with either HP-1 or HP-5 stationary phase (film thickness of (Frimmel et al., 2004) occur at ∼6.5 m and 12.94 m above the base of the
0.25 µm). The carrier gas was helium, and for analysis of the aliphatic formation and coincide with the stepped increases in %TOC suggesting
hydrocarbons (see Fig. 4) the oven temperature was held at 40 °C for major contributions to the organic matter from plankton (Fig. 6C).
2 min and then heated at a rate of 4 °C/min to 300 °C at which it was The mass chromatogram for m/z 133 from the aromatic hydrocarbon
held for 20 min. The following oven temperature programme was fraction (Fig. 7) reveals a pseudo-homologous series of aryl isoprenoids
used for the analysis of the aromatic hydrocarbons (primarily for the up to C26 (see I through to X in Fig. 7 and Table 2) as well as the presence
assignment of isorenieratane XXIII): the oven was held at 60 °C for of aryl isoprenoids with additional aromatic rings (see XI and XVI in
2 min and then heated to 240 °C at 10 °C/min, further heated to 315 °C Fig. 7 and Table 2). The relative amounts of the different components
at 4 °C/min where it was held at the final temperature for 50 min. remains the same throughout the profile where the C17, C20, C21, C22

Fig. 4. Total ion chromatogram (TIC) of aliphatic hydrocarbon fraction from Upper Ordovician black shale from 16.76 m (sample 8402-77) above the base of the formation in BG-14
borehole. Pr = pristane, Ph = phytane; numbers denote total number of carbon atoms in the n-alkanes.
372 H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377

Fig. 5. GC-MS ion chromatogram of m/z 133 of aromatic hydrocarbon fraction from Upper Ordovician black shale (sample 3402-15) taken at 33.9 m in the core, 12.94 m above the base
of the formation. (A) after and (B) before co-injection with an authentic standard of isorenieratane (XXIII; courtesy of S. Schouten, Netherlands Institute for Sea Research (NIOZ), Den
Burg, Netherlands) showing enhancement of peak after co-injection. (C) is mass spectrum of isorenieratane (XXIII) from the sample.

members have reduced abundances. This distribution is similar to co-elution with an authentic standard on a range of stationary phases
that observed in Upper Devonian sediments (Hartgers et al., 1994) but (Sinninghe Damsté et al., 2001).
differs from that in much older mid-Proterozoic bitumens (Brocks
et al., 2005). There are also less abundant C32, C33 and C40 diaryl 5. Discussion
isoprenoids both with and without an additional aromatic ring (see
XVII through to XXIII in Fig. 7 and Table 2). Isorenieratane (XXIII) was The deposition of sedimentary organic carbon in the lower Batra
present throughout the core and its identity was also confirmed by Formation is delimited by stepped increases (approximate doubling at
H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377 373

Fig. 6. Composite plot of bulk organic carbon, biomarker and bulk stable carbon isotopic data. (A) Total organic carbon (TOC) content of the bulk sediment. (B) Hydrogen index (HI) of
the bulk sediment (mg hydrocarbons (HC)/g TOC). (C) Steranes/17α-hopanes ratio shows its highest value at 12.94 m above the base of the Batra formation. (D) δ13C values of organic
carbon (OC) versus Vienna Peedee belemnite (VPDB) in parts per mil (‰).

each step) in %TOC up succession. %TOC has been found to correlate (Fig. 6B and D); and this with the unchanging nature of kerogen type
well with more direct measures of photosynthetic primary produc- through the section, suggests variations in HI reflect the changing extent
tivity such as total chlorophyll-a or total steryl chlorine esters (Nara of organic matter preservation (see Tyson, 1995).
et al., 2005) and organic mass accumulation rate (Tyson, 1995; Peaks in steranes/17α-hopanes ratio (N1) coincide with the stepped
Vilinski and Domack, 1998; Twichell et al., 2002; Meyers and changes in %TOC, that provide a coherent signal of primary productivity.
Arnaboldi, 2005). A longer-term trend towards more fractionated We therefore conclude the changes in %TOC reflect changes in plankton
values up succession reflects a progressive increase in the sedimen- primary productivity in the photic zone. The decay of this organic matter
tation of 12C-enriched organic matter. through oxidative respiration resulted in anoxia and euxinia, which
Hydrogen index (HI) is a measure of hydrogen-richness and depends enhanced the potential for organic matter preservation.
on the nature of the original organic matter and the degree of Our bulk δ13C values fall within the range for modern phytoplanktonic
preservation during diagenesis (Peters, 1986; Bordenave et al., 1993). algae (Schidlowski, 1988) and for bulk organic matter in the Southern
Millimetre-scale laminations and presence of macroscopic pyrite Ocean at 0 °C (Rau et al., 1989; Bentaleb and Fontugne, 1998; Bentaleb
through the section suggests euxinia and that organic matter preserva- et al., 1998; Lourey et al., 2004). Increasingly more fractionated bulk
tion potential was high throughout deposition. Our bulk rock HI values organic matter δ13C values occur up section with no change in kerogen
are similar to those reported from Mediterranean, Pliocene–Pleistocene type. In ice margin basins at the present day changes in δ13C are usually
(b3 myr old) immature sapropels which have HI values of ∼400 mg HC/g associated with a decreased CO2 availability in response to (a) decreased
TOC (Tyson, 1995). Bulk HI and organic matter δ13C values do not covary supply by diffusive limitation, or (b) increased demand because of higher

Fig. 7. Typical GC-MS summed mass chromatogram of m/z 133 + 134 from the aromatic hydrocarbon fraction isolated from the shale organic extract at 11.64 m above the base of the
Batra Formation. This trace shows the C40 biomarker isorenieratane (XXIII). Roman numerals refer to compounds indicated in Appendix A and Table 2.
374 H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377

Table levels
light 2 and growth rates (Laws et al.,1995; McMinn et al.,1999) and/or stratification is maintained in these basins by an increase in salinity
Details of mass spectral
the increased identifications
availability of I–XXIII
of (trace) nutrients at the time of plankton (Gibson, 1999) resulting from brine exclusion during sea ice formation.
13
growth
Peak label(Redfield et al.,
Molecular 1963). Molecular
formula Severe fractionation
ion(m/z) of δfragment
Major C is known to
ions (m/z) During the winter, a thermocline convection cell develops directly
I C13H20 176 133/134 beneath the ice cover and penetrates progressively deeper into the
II C14H22 190 133/134 basin throughout winter. At the end of the period of ice formation the
III C15H24 204 133/134 convection cell breaks down and stratification of the surface water
IV C16H26 218 133/134
occurs. When the ice melts completely, lenses of relatively fresh
V C17H28 232 133/134
VI C18H30 246 133/134 water cap the basins; this reduces the effect of wind mixing, with a
VII C19H32 260 133/134 net result of stabilising the basin stratification, with anoxia develop-
VIII C20H34 274 133/134 ing at depth. The effect of, increasing water level in the basins or
IX ? 226 133/134, 147, 220 decreasing maximum ice thickness during the summer results in a
X C21H36 288 133/134, 220
XI C21H27 280 133/134
shallowing of the mixocline and chemocline (Gibson, 1999). The
XII ? 302 133/134, 185, 253 nature of the stratification in these basins (Gibson, 1999; McMinn
XIII C22H29 294 133/134 et al., 2001) is therefore similar to those reported from the many
XIV ? 336 133/134, 239, 253 permanently stratified basins around the world including the Red Sea
XV ? 350 133/134, 195, 239, 253
(Hartmann et al., 1998), the Cariaco Trench, the Black Sea, and fjords
XVI C26H32 344 133/134, 169/170, 253
XVII C32H42 426 133/134, 119, 173 of Scandinavia (Skei, 1983; Lindholm, 1996).
XVIII C32H50 434 133/134 We envisage the geological history of the Batra Basin and its included
XIX C33H42 448 133/134 sediments to be directly controlled by ice margin processes. On ice
XIX C33H42 448 133/134 melting and retreat the exposed upper palaeovalley became a conduit for
XX C40H58 538 133/134, 237
ice meltwater. The basin was initially isolated from shallow marine
XXI C40H58 538 133/134, 173
XXII C40H66 546 133/134, 235 waters, likely silled and glaciofluvial sediments were deposited. As the
XXIII C40H66 546 133/134 effects of isostatic rebound waned and sea level rose, the basin was
flushed by marine waters. On short timescales we envisage the basin
became stratified due to the formation and melting of sea ice (Fig. 8). The
millimetre-scale laminations in the black shales may reflect seasonal to
occur naturally in either intense phytoplankton blooms (Dunbar and decadal (? millennial) changes in sea ice cover. On the longer term,
Leventer, 1992) or within sea ice (Dunbar and Leventer, 1992; McMinn periods of increased surface primary productivity occurred when
et al.,1999). Natural values for modern Southern Ocean phytoplankton can prolonged ice free conditions prevailed and/or, fluxes of freshwater and
be as low as −25‰ (Rau et al., 1991). While highly fractionated values as nutrients entered the already stratified, euxinic marine basin. Stratigra-
high as −12‰ have been recorded from sea ice (Dunbar and Leventer, phical evidence (Armstrong et al., 2005) is consistent with modelling
1992; McMinn et al., 1999). Thus when sea ice melts it not only delivers a results (Herrmann et al., 2003) and indicates the frequency of melting
pulse of sediment and nutrients, but also organic matter that has a more events during the late Hirnantian deglaciation was likely to have occurred
fractionated δ13C signature (Gibson, 1999). on an obliquity (∼40 kyr) timescale. The black shales of the Batra Basin
A pattern of long-term covariance of more fractionated δ13C and record a few hundred thousand years of water column stratification and
increasing %TOC has also been reported in Cretaceous (Late Cenoma- basin euxinia during deglacial highstand. This represents one of the
nian/Turonian) black shale in the North Atlantic, deposited during oldest peri-glacial permanently stratified basins yet described.
periods of enhanced continental run-off (Beckmann et al., 2005a,b).
Here the pattern has been interpreted to indicate organic carbon 6. Conclusions
sequestration to the sediment was cumulative through the deposi-
tional event (Kuypers et al., 2002). Sedimentology, geological setting and organic geochemical
We therefore consider the more fractionated δ13C values through proxy data indicate the Batra Formation black shale in Jordan was
the lower Batra section to represent CO2 limitation and increased deposited in a permanently stratified, ice margin, marine basin that
productivity as nutrients and isotopically light carbon were supplied existed for a few hundred thousand years. Euxinia extended into the
by melt water. The long-term trend towards more fractionated values photic zone enhancing sedimentary carbon preservation and
reflects a progressive increase in the sedimentation of 12C-enriched sedimentation. Ice melting and/or, fluxes of freshwater and
organic matter during progressive deglaciation. nutrients resulted in enhanced photic zone primary productivity
The presence of isorenieratene derivatives in black shales has been
widely considered diagnostic of green sulphur reducing bacteria
(Chlorobiaceae and Chromatiaceae) and is used as evidence for photic
zone euxinia (e.g. Sinninghe Damsté and Köster, 1998). Further, pyrite
is common through the section. In the modern ocean (Killops and
Killops, 1993) the depth of the sulphate reduction zone depends on
the amount of organic matter influx from the euphotic zone and may
be relatively shallow (b20 m) in highly productive areas, where
sulphate is rapidly depleted (Brocks et al., 2005). The depth to the
chemocline in the Batra Basin may have been shallower than the
∼ 50 m found in the Black Sea at the present day (Murray et al., 1989).
The greatest concentration of stratified water bodies in high
latitudes, and possibly the world, is found in the Vestfold Hills of
Antarctica. Here meromictic lakes, isolated marine basins and fjords
occur (Burton,1981; Burke and Burton,1988a; Gallagher et al., 1989) and
seasonal anoxia is developed in these settings. These basins formed
following the retreat of the continental ice sheet ∼ 10 000 years ago, Fig. 8. Conceptual model of the open water (ice free), stratified water column during the
when isostatic rebound occurred at a faster rate than sea level rise, and deposition of the lower “hot shales” in the Batra Basin. This is based in part on that found
the land emerged from the sea (Burke and Burton, 1988b). Seasonal in modern ice margin basins in the Vestfold Hills, east Antarctica (Gibson, 1999, Fig. 7).
H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377 375

and organic matter sedimentation. The seasonally isolated basins and JREI awards. Dr David Loydell (University of Portsmouth) and Dr Mark
anoxic fjords of east Antarctica provide a likely modern analogue, Williams (University of Leicester) kindly identified the graptolites. TOC
though these are b10 000 years old. Black shale was patchily measurements were conducted by Dr D.M. Jones (University of
deposited along the entire northern Gondwana margin during the Newcastle upon Tyne) and isotopic analyses by Prof. M. Leng (NERC
Silurian. Similarly detailed interdisciplinary studies are required to Isotope Geosciences Laboratory). We thank Christine Jeans for the
test whether coastal upwelling can be invoked to explain the origin of preparation of the Figures and B. Bowler for technical input. H.P.
any of these deposits. acknowledges the European Commission Research Directorates Gen-
eral for a Marie Curie Host Fellowship held at the University of
Acknowledgements Newcastle upon Tyne. A.B.M. was supported by the Petroleum
Technology Development Fund, Nigeria. H.A.A. and B.R.T. acknowledge
The National Resources Authority of Jordan provided access to the funding from the Natural Environment Research Council. This is a
core. We acknowledge support from our host institutions and the contribution to IGCP Project 503. We thank Lorenz Schwark and an
Natural Environment Research Council for research funds including anonymous referee for their helpful suggestions.

Appendix A. Structures of isorenieratene derivatives


376 H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377

References Laws, E.A., Popp, B.N., Bidigare, R.R., Kennicutt, M.C., Macko, S.A., 1995. Dependence of
phytoplankton carbon isotopic composition on growth rate and [CO2(aq)].
Theoretical considerations and experimental results. Geochimica et Cosmochimica
Abed, A.M., Makhlouf, I.M., Amireh, B.S., Khalil, B., 1993. Upper Ordovician glacial Acta 59, 1131–1138.
deposits in southern Jordan. Episodes 16, 316–328. Lindholm, T., 1996. Periodic anoxia in an emerging coastline basin in SW Finland.
Amireh, B.S., Schneider, W., Abed, A.M., 2001. Fluvial-shallow marine-glaciofluvial Hydrobiologia 325, 223–230.
depositional environments of the Ordovician System in Jordan. Journal of Asian Lourey, M.J., Thomas, W., Trull, T., Tilbrook, B., 2004. Sensitivity of d13C of Southern
Earth Sciences 19, 45–60. Ocean suspended and sinking organic matter to temperature, nutrient utilization
Andrews, I.J., 1991. Palaeozoic lithostratigraphy in the subsurface of Jordan. Subsurface and atmospheric CO2. Deep-Sea Research I 51, 281–305.
Geology Bulletin 2. Loydell, D.K., 1998. Early Silurian sea-level changes. Geological Magazine 135, 447–471.
Armstrong, H.A., 2007. On the cause of the Ordovician glaciation. In: Williams, M., Loydell, D.K., 2007. Graptolites from the Upper Ordovician and Lower Silurian of Jordan.
Haywood, A., Gregory, J. (Eds.), Deep Time Perspectives on Climate Change. Geological Special Papers in Palaeontology, vol. 78.
Society of London. The Micropalaeontological Society, London, pp. 101–121. Lüning, S., Craig, J., Loydell, D.K., Storch, P., Fitches, B., 2000. Lower Silurian ‘hot shales’ in
Armstrong, H.A., Turner, B.R., Makhlouf, I.A., Williams, M., Al Smadi, A., Abu Salah, A., North Africa and Arabia: regional distribution and depositional model. Earth
2005. Origin, sequence stratigraphy and depositional environment of an Upper Science Reviews 49, 121–200.
Ordovician (Hirnantian), peri-glacial black shale, Jordan. Palaeogeography, Palaeo- Lüning, S., Loydell, D.K., Storch, P., Shahin, Y., Craig, J., 2006. Origin sequence
climatology, Palaeoecology 220, 273–289. stratigraphy and depositional environment of an upper Ordovician (Hirnantian)
Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A., deglacial black shale, Jordan. Discussion. Palaeogeography, Palaeoclimatology,
Abu Salah, A., 2006. Reply to “Origin, sequence stratigraphy and depositional Palaeoecology 230, 352–355.
environment of an upper Ordovician (Hirnantian) deglacial black shale, Jordan”. Makhlouf, I.M., 1995. Tempestite facies displaying hummocky cross-stratification and
Palaeogeography, Palaeoclimatology, Palaeoecology 230, 356–360. subaqueous channels in Ordovician shelf deposits, South Jordan. Africa Geosciences
Beckmann, B., Flögel, S., Hofmann, P., Schulz, M., Wagner, T., 2005a. Orbital forcing of Review 2, 91–99.
Cretaceous river discharge in tropical Africa and ocean response. Nature 437 (8), Masri, A., 1988. The Geology of Halat Ammar and Al Mudawwara, Map Sheet Nos. 3248
241–244. III, 3248 IV, 13. Natural Resources Authority, Geological Mapping Directorate,
Beckmann, B., Wagner, T., Hofmann, P., 2005b. Linking Coniacian–Santonian (OAE3) Geological Mapping Division, Amman, Jordan, 59 pp.
black shale deposition to African climate variability: a reference section from the McMinn, A., Skerratt, J.H., Trull, T., Ashworth, C., Lizotte, M., 1999. Nutrient stress
eastern tropical Atlantic at orbital timescales (ODP site 959, Off Ivory Coast and gradient in the bottom 5 cm of fast ice, McMurdo Sound, Antarctica. Polar Biology 21,
Ghana). SEPM Special Publication 82, 125–143. 220–227.
Bentaleb, I., Fontugne, M., 1998. The role of the Southern Indian Ocean in the glacial to McMinn, A., Heijnis, H., Harle, K., McOrist, G., 2001. Late Holocene climatic change
interglacial atmospheric CO2 change: organic carbon isotope evidence. Global and recorded in sediment cores from Ellis Fjord, eastern Antarctica. The Holocene 11,
Planetary Change 16–17, 25–36. 291–300.
Bentaleb, I., Fontugne, M., Descolas-Gros, C., Girardin, C., Mariotti, A., Pierre, C., Brunet, Meyers, P.A., Arnaboldi, M., 2005. Trans-Mediterranean comparison of geochemical
C., Poisson, A., 1998. Carbon isotopic fractionation by phytoplankton in the Southern productivity proxies in a mid-Pleistocene interrupted sapropel. Palaeogeography,
Indian Ocean: relationship between d13C of particulate organic carbon and Palaeoclimatology, Palaeoecology 222, 313–328.
dissolved carbon dioxide. Journal of Marine Systems 17, 39–58. Miller, M.A., Mansour, H.A.-R., 2007. Preliminary palynological investigation of Saudi
Bordenave, M.L., Espitalié, L., Leplat, P., Oudin, J.L., Vandenbroucke, M., 1993. Screening Arabian Upper Ordovician glacial sediments. Revue de Micropaleontologie 50, 17–26.
techniques for source rock evaluation. In: Bordenave, M.L. (Ed.), Applied Petroleum Murray, J.W., Jannasch, H.W., Honjo, S., Anderson, R.F., Reeburgh, W.S., Top, Z.,
Geochemistry. Editions Technip, Paris, pp. 217–278. Friederich, G.E., Codispoti, L.A., Izdar, E., 1989. Unexpected changes in the oxic/
Brocks, J.J., Love, G.D., Summons, R.E., Knoll, A.H., Logan, G., Bowden, S.A., 2005. anoxic interface in the Black Sea. Nature 338, 411–414.
Biomarker evidence for green and purple sulfur bacteria in an intensely stratified Nara, F., Tani, Y., Soma, Y., Soma, M., Naraoka, H., Watanabe, T., Horiuchi, K., Kawai, T.,
Paleoproterozoic sea. Nature 437, 866–870. Oda, T., Nakamura, T., 2005. Response of phytoplankton productivity to climate
Burton, H.R., 1981. Chemistry, physics and evolution of Antarctic saline lakes. change recorded by sedimentary photosynthetic pigments in Lake Hovsgol
Hydrobiologia 82, 339–362. (Mongolia) for the last 23,000 years. Quaternary International 136, 71–81.
Burke, C.M., Burton, H.R., 1988a. The ecology of photosynthetic bacteria in Burton Lake, Page, A.A., Zalasiewicz, J.A., Williams, M., Popov, L.E., 2007. Were transgressive black
Vestfold Hills, Antarctica. Hydrobiologia 165, 1–11. shales a negative feedback modulating glacioeustacy in the Early Palaeozoic
Burke, C.M., Burton, H.R., 1988b. Photosynthetic bacteria in meromictic lakes and icehouse? In: Williams, M., Haywood, A., Gregory, F.J., Schnmidt, D.N. (Eds.), Deep-
stratified fjords of the Vestfold Hills, Antarctica. Hydrobiologia 165, 13–23. Time perspectives on Climate Change: Marrying the Signal from Computer Models
Cocks, L.R.M., Rickards, R.B., 1988. A global analysis of the Ordovician–Silurian and Biological Proxies. Special Publication of the Geological Society of London. The
boundary. Special Issue of the Bulletin of the British Museum (Natural History). Micropalaeontological Society, pp. 123–156.
Geology 43 394 pp. Peters, K.E., 1986. Guidelines for evaluating petroleum source rock using programmed
Cocks, L.R.M., Torsvik, T.H., 2002. Earth geography from 500 to 400 million years ago: a pyrolysis. Bulletin of the American Association of Petroleum Geologists 70, 318–329.
faunal and palaeomagnetic review. Journal of the Geological Society of London 159, Peters, J.M., Walters, C.C., Moldowan, J.M., 2005. The Biomarker Guide: Biomarkers and
631–644. Isotopes in the Environment and Human History. Cambridge University Press,
Droser, M.L., Bottjer, D.J., 1986. A semiquantitative field classification of ichnofabric. Cambridge. 471 pp.
Journal of Sedimentary Petrology 56, 558–559. Powell, J.H., Moh'd, B.K., Masri, A., 1994. Late Ordovician–Early Silurian glaciofluvial deposits
Dunbar, R.B., Leventer, A., 1992. Seasonal variation in carbon isotopic composition of preserved in palaeovalleys in South Jordan. Sedimentary Geology 89, 303–314.
Antarctic sea ice and open water plankton communities. Antarctic Journal of the US Rau, G.H., Takahashi, T., Des Marais, D.J., 1989. Latitudinal variations in plankton d13C:
27, 79–81. implications for CO2 and productivity in past oceans. Nature 341, 516–518.
Frimmel, A.W., Oschmann, W., Schwark, L., 2004. Chemostratigraphy of the Posidonia Rau, G.H., Froelich, P.N., Takahashi, T., Des Marais, D.J., 1991. Does sedimentary organic
Black Shale, SW Germany. I. Influence of sea level variation on organic facies d13C record variations in Quaternary ocean [CO2(aq)]? Paleoceanograpy 6, 335–347.
evolution. Chemical Geology 206, 199–230. Redfield, A.C., Ketchum, B., Richards, F.A., 1963. The influence of organisms on the
Gallagher, J.B., Burton, H.R., Calf, G.E., 1989. Meromixis in an Antarctic Fjord: a precursor composition of seawater. In: Hill, M.N. (Ed.), The Sea, vol. 2. Wiley-Interscience,
to meromictic lakes on an isostatically rising coastline. Hydrobiologia 172, 235–254. New York, pp. 26–77.
Gibson, J.A.E., 1999. The meromictic lakes and stratified marine basins of the Vestfold Schidlowski, M., 1988. A 3800-million year isotopic record of life from carbon and
Hills, East Antarctica. Antarctic Science 11, 175–192. sedimentary rocks. Nature 333, 313–318.
Grantham, P.J., Wakefield, L.L., 1988. Variations in the steranes carbon number Sinninghe Damsté, J.S., Köster, J., 1998. A euxinic southern North Atlantic Ocean during
distributions of marine source rock derived crude oils through geological time. the Cenomanian–Turonian oceanic anoxic event. Earth and Planetary Science
Organic Geochemistry 12, 61–73. Letters 158, 165.
Hartgers, W.A., Sinninghe Damsté, J.S., Requejo, A.G., Allan, J., Hayes, J.M., Ling, Y., Xie, T.-M., Sinninghe Damsté, J.S., Schouten, S., van Duin, A.C.T., 2001. Isorenieratene derivatives in
Primack, J., de Leeuw, J.W., 1994. A molecular and carbon isotopic study towards the sediments: possible controls on their distribution. Geochimica et Cosmochimica
origin and diagenetic fate of diaromatic carotenoids. Organic Geochemistry 22, 703–725. Acta 65, 1557–1571.
Hartmann, M., Scholten, J.C., Stoffers, P., Wehner, F., 1998. Hydrographic structure of Skei, J., 1983. Geochemical and sedimentological considerations of a permanently
brine-filled deeps in the Red Sea—new results from the Shaban, Kebrit, Atlantis II anoxic fjord—Framvaren Fjord, South Norway. Sedimentary Geology 36, 131–145.
and Discovery Deep. Marine Geology 144, 311–330. Storch, P., 1990. Upper Ordovician–lower Silurian sequences of the Bohemian Massif,
Herrmann, A.D., Patzkowsky, M.E., Pollard, D., 2003. Obliquity forcing with 8–12 times central Europe. Geological Magazine 127, 225–239.
preindustrial levels of atmospheric pCO2 during the Late Ordovician glaciation. Sutcliffe, O.E., Dowdeswell, J.A., Whittington, R.J., Theron, J.N., Craig, J., 2000. Calibrating
Geology 31, 485–488. the Late Ordovician glaciation and mass extinction by the eccentricity of Earth's
Keegan, J.B., Rasul, S.M., Shaheen, Y., 1990. Palynostratigraphy of the Lower Paleozoic, orbit. Geology 28, 967–970.
Cambrian to Silurian, sediments of the Hashemite Kingdom of Jordan. Review of Turner, B.R., Armstrong, H.A., Makhlouf, I.M., Bourne, T.J., 2002. High latitude, east
Palaeobotany and Palynology 66 (3–4), 167–180. Gondwana glaciation: glacio-fluvial palaeovalleys interpreted as tunnel valleys.
Killops, S.D., Killops, V.J., 1993. An Introduction to Organic Geochemistry. Longman Gondwana 11 Programme with Abstracts.
Scientific and Technical, Harlow. 265 pp. Turner, B.R., Makhlouf, I.M., Armstrong, H.A., 2005. Late Ordovician (Ashgillian) glacial
Kuypers, M.M.M., Pancost, R.D., Nijenhuis, I.A., Sinninghe Damsté, J.S., 2002. Enhanced deposits in southern Jordan. Sedimentary Geology 181, 73–91.
productivity led to increased organic carbon burial in the euxinic North Atlantic Twichell, S.C., Meyers, P.A., Diester-Haass, L., 2002. Significance of high C/N ratios in
basin during the late Cenomanian oceanic anoxic event. Paleoceanography 17, 1051. organic carbon-rich Neogene sediments under the Benguela Current upwelling
doi:10.1029/2000PA000569. system. Organic Geochemistry 33, 715–722.
H.A. Armstrong et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 368–377 377

Tyson, R.V., 1995. Sedimentary Organic Matter: Organic Facies and Palynofacies. Wignall, P.B., 1991. Model for transgressive black shales? Geology 19, 167–170.
Chapman and Hall, London. 615 pp. Wignall, P.B., 1994. Black Shales. Clarendon Press, Oxford. 127 pp.
Vaslet, D., 1990. Upper Ordovician glacial deposits in Saudi Arabia. Episodes 13, 147–161. Zalasiewicz, J.A., Tunnicliff, S.P., 1994. Uppermost Ordovician to lower Silurian graptolite
Vilinski, J.C., Domack, E., 1998. Temporal changes in sedimentary organic carbon from biostratigraphy of the Wye Valley, central Wales. Palaeontology 37 (3), 695–720.
the Ross Sea Antarctica: inferred changes in ecosystems and climate. Eos
(Transactions, American Geophysical Union) 79, 157.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s e v i e r. c o m / l o c a t e / p a l a e o

Palynology, organic geochemistry and carbon isotope analysis of a latest Ordovician


through Silurian clastic succession from borehole Tt1, Ghadamis Basin, southern
Tunisia, North Africa: Palaeoenvironmental interpretation
Marco Vecoli a,⁎, Armelle Riboulleau b, Gerard J.M. Versteegh c
a
Research team UMR 8157, CNRS, University of Lille 1, Villeneuve d'Ascq, F-59655, France
b
Research team UMR 8157, PBDS, University of Lille 1, Bâtiment SN5, 59655 Villeneuve d'Ascq, F-59655, France
c
Organic Geochemistry Unit, Faculty of Geosciences, Bremen University, P.O. Box 330440 D-28334 Bremen, Germany

A R T I C L E I N F O A B S T R A C T

Article history: Palynological and palynofacies analyses combined with organic carbon isotope measurements have been
Received 12 December 2007 performed on terminal Ordovician through Silurian clastic sediments from the North African margin of
Received in revised form 26 May 2008 Gondwana (southern Tunisia). A first carbon isotopic signal (δ13Corg) from Gondwanan Silurian sedimentary
Accepted 27 May 2008
sequences is presented, showing interesting correlation with existing coeval isotopic curves from other areas.
Changes in lithology, palynofacies characteristics, palynomorph diversity, carbon isotope developments, and
Keywords:
Palynomorphs
organic geochemistry parameters appear to be all causally linked, and to reflect changes in palaeoceano-
Palynofacies graphic conditions. In particular, the detailed chronostratigraphic correlation and the observed similarities in
Carbon isotopes patterns of carbon isotopic developments and (palyno- and litho-) facies changes through the study section
Bioevents has permitted to identify the supposedly global earliest Wenlock (“Ireviken Event”) and late Ludlow (“Lau
Black shales Event”) isotopic excursions for the first time in high-latitude Gondwana. This confirms that these excursions
Productivity were linked to global changes in the oceanic system.
Gondwana The present results suggest that from Rhuddanian to early Wenlock times, an extended period of black shale
Ordovician
deposition occurred over the North African Gondwanan margin, progressively transgressing from basin
Silurian
palaeodepressions to basin palaeohighs. Palynofacies and organic geochemistry support a coastal upwelling-
promoted productivity increase during this interval, associated to a decrease in diversity of the marine
microplanktonic communities. The earliest Wenlock strong positive isotopic shift appears associated with
this protracted period of massive black shale deposition, and thus of organic carbon burial, on continental
platforms located in high-latitude settings. This could well explain the apparent paradox between excessive
carbonate deposition simultaneous to a carbon isotopic shift towards lighter values observed in low
palaeolatitude localities (Laurentia and Baltica). the strong, distinct isotope excursion occurring in late
Ludlow times is possibly linked to the well know Lau Event. Evidence for extensive organic carbon burial is
lacking to explain this strong excursion. Some significant changes in the marine palynomorph communities
are recorded in connection to the increase in stable isotope values (δ13Corg). Additionally, a strong correlation
between abundance of terrestrially derived palynomorphs (miospores) and δ13Corg development is recorded
which seems to support increased clastic input from a terrestrial source.
© 2008 Elsevier B.V. All rights reserved.

1. Introduction black shales with exceptionally high organic content (“hot shales”:
Lüning et al., 2000), and the progressive establishment of a well
Ordovician-Silurian sedimentary sequences at the northern mar- developed vegetation cover on the continents.
gin of western Gondwana record palaeoclimatic and palaeoenviron- Previous studies have documented these latest Ordovician-Silurian
mental changes of global significance. These include the end of the events in various areas of the planet, progressively evidencing
severe and short-lived Late Ordovician glaciation which caused one of causative links between changes in the geosphere (variations in
the major faunal extinctions of the Phanerozoic, the successive return global climate, carbon cycling, sea-level, surface-rock weathering,
to pre-glacial, globally warmer conditions, widespread deposition of sediment fluxes: e.g., Kaljo et al., 1988; Samtleben et al., 2000; Cramer
and Saltzman, 2005, 2007a,b) and perturbations in the biosphere
(faunal and floral extinctions, radiations, rates of biodiversity changes:
⁎ Corresponding author. Jeppsson, 1987, 1990; Kaljo et al., 1995; Stricanne et al., 2006). This has
E-mail address: Marco.Vecoli@univ-lille1.fr (M. Vecoli). shown that the Silurian, once considered as a period of relatively

0031-0182/$ – see front matter © 2008 Elsevier B.V. All rights reserved.
doi:10.1016/j.palaeo.2008.05.015
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 379

stable palaeoenvironmental conditions, was instead punctuated by coupled to detailed analysis of extinction patterns and fossil
major instabilities in palaeoceanography, climate, carbon cycling on biodiversity dynamics, have proved to be very useful for palaeoenvi-
time scales longer than 105 years (Samtleben et al., 2000; Munnecke ronmental reconstructions. A series of distinct excursions in oxygen
et al., 2003), as well as by extinction events in different fossil groups and carbon isotope values have been described (e.g., Munnecke et al.,
(Kaljo et al., 1995; Munnecke et al., 2003). Isotopic investigations 2003), which coincide with distinct lithological and biotic changes,

Fig. 1. Geographic location of borehole Tt1 in the context of main geological features of the Sahara Platform. Modified from Storch and Massa (2007).
380 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

and seem to have global significance. However, most of the data Beuf et al., 1971; Legrand, 1985). Palaeogeographically, North Africa
available on Silurian stable isotopic developments and patterns of was located at the margin of Gondwana in high southern palaeola-
extinctions comes principally from Baltican and Laurentian localities titudes (Fig. 2). The Lower Palaeozoic sedimentary successions of the
(Saltzman, 2002; Munnecke et al., 2003; Martma et al., 2005; Kaljo Sahara Platform are mainly detrital in nature, have not been affected
and Martma, 2006; Cramer and Saltzman, 2007a,b). by metamorphism, and generally contain well preserved, often
The well preserved sedimentary sequences occurring in the thermally unaltered organic matter, including exceptionally well
subsurface of North Africa represent an appropriate target in order preserved palynomorphs (acritarchs, chitinozoans, miospores; Jardiné
to provide the hitherto missing Gondwanan perspective of latest et al., 1974; El Arnauti et al., 1988; Paris, 1990; Vecoli, 2000). The
Ordovician through Silurian environmental change. Ghadamis Basin is mainly developed in northwestern Libya, continu-
In this paper, palynofacies and palynostratigraphic investigations ing into southern Tunisia and eastern Algeria (Fig. 1), covering an area
are combined with organic carbon isotope measurements of terminal of approximately 200.000 km2 (Massa, 1988; Rusk, 2001). The study
Ordovician through Silurian predominantly shaly sediments from the area, were borehole Tt1 is located, corresponds to a structural high
northern Ghadamis Basin in the subsurface of southern Tunisia. We delimiting the northernmost extension of the Ghadamis Basin, known
provide the first carbon isotopic signal from Gondwanan Silurian as the Sidi-Toui uplift (“Môle de Sidi-Toui”; Massa, 1988, p. 39, Fig. 9),
sedimentary sequences and couple it to coeval isotopic curves from which is the western continuation of the conspicuous Nefusah Uplift
other areas. This enables us to set precise chronostratigraphic of northeastern Libya (Figs. 1 and 2; Massa, 1988).
constraints based on detailed fossil occurrences on the earliest to
early Silurian black shale units (e.g. “hot shales”) which has been a 3. Borehole stratigraphy
somewhat controversial issue (cf., Lüning et al., 2000; Armstrong
et al., 2005, 2006; Lüning et al., 2006). In combination with the The investigated core section ranges between −1330 m and −
observed palynofacies developments and changes in palynomorph 1225 m core depth in borehole Tt1, spanning the latest Ordovician–
(mainly marine microphytoplankton and miospores) diversity we late Silurian (Ludlow) stratigraphic interval (Figs. 3 and 4). The
shed more light on the origin of the Silurian black shales and the lithostratigraphic terminology adopted here follows the one defined
environmental changes involved. The insights obtained contribute to by Massa and Jaeger (1971), Jaeger et al. (1975), and Massa (1988) for
our global understanding of the evolution of life and its environment the Ghadamis Basin.
during the Silurian. The lower part of the section, between ca. −1330 m and ca. −
1302 m consists of irregular alternations of fine-grained, immature
2. Geological setting sandstones, siltstones and dark grey shales characterized by abundant
isolated angular to subangular quartz grains. These sediments are
Study material comes from a subsurface section in southern considered to be of distal, peri-glacial origin and are attributed to the
Tunisia, in proximity of the Libyan border (borehole Tt1, Ghadamis Djeffara Formation (Jaeger et al., 1975; Massa, 1988). Recent
Basin; Fig. 1). This area is a part of the vast sedimentary zone which palynostratigraphic investigations of the widespread Ordovician,
crosses northern Africa between parallels 24 and 36° N lat, from the glacial-related sediments over the entire Sahara Platform (e.g.,
Atlantic Ocean to the Red Sea, also known as the Saharan, or North “Argiles Microconglomératiques” in Algeria, the Djeffara Fm. in
African, Platform. Vast and laterally continuous subcrops of Palaeozoic Tunisia/NW Libya, the “2nd Bani” Fm. in Morocco), demonstrate an
sedimentary successions are known to occur over the greater part of Hirnantian age (latest Ordovician; Paris et al., 1995, 2000; Bourahrouh
the Sahara Platform (Beuf et al., 1971). They are mainly developed into et al., 2004; Ghienne et al., 2007; Vecoli, 2008). Three samples were
large intracratonic basins, delimited by broadly elongated positive collected from this interval (Figs. 3 and 4).
structures showing a general north–south trend, and swinging Above the Djeffara Formation, a suite of fine grained, predomi-
westward or eastward to the north (Balducchi and Pommier, 1970; nantly shaly to silty sediments with marly interbeds occurs. This

Fig. 2. Early Silurian palaeogeographic reconstruction (based on Scotese and McKerrow, 1990) with position of study area (open circle) and other localities documenting Silurian
isotopic events, as discussed in text. Numbers on map refer to several reports and multiple stratigraphic sections: 1 — North America (Laurentia; Saltzman, 2001; Cramer and
Saltzman, 2005; Cramer et al., 2006; Cramer and Saltzman, 2007a); 2 — Sweden (Baltica; Samtleben et al., 2000; Munnecke et al., 2003; Stricanne et al., 2006); 3 — Lithuania and
Estonia (Baltica; Martma et al., 2005; Kaljo and Martma, 2006); 4 — Australia (eastern Gondwana; Talent et al., 1993; Andrew et al., 1994); 5 — Prague Basin, Czech republic (peri-
Gondwana; Lehnert et al., 2007).
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394
381
Fig. 3. Lithostratigraphy, biostratigraphy, chronostratigraphic attributions, sampling levels in borehole Tt1, and stratigraphic occurrences of microphytoplankton species (acritarchs and prasinophytes).
382 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

sedimentary succession is particularly rich in graptolites, which of the “vesiculosus” zone (Rhuddanian, early Llandovery: Jaeger et al.,
permit reliable dating with a high degree of biostratigraphic precision, 1975 and Storch, pers. comm., September 2007). From this interval,
indicating an age range from the early Llandovery to the Ludlow five samples were available for analysis (Figs. 3 and 4).
(Jaeger et al., 1975; Massa, 1988). Based on several subcrop sections in Between −1298 m and −1267 m (thirteen samples collected; Figs. 3
southern Tunisia and northwestern Libya, Jaeger et al. (1975) formally and 4) a distinct interval occurs which is mainly composed of dark
named this early to late Silurian, silty–shaly succession as the “Argiles grey to black shales and marls, rarely interbedded with limestone/
Principales” Formation; distinguishing it from the broadly coeval but dolomite layers. Organic-rich black shales and shaly marls occur in
clearly more proximal Tanezzuft Formation as defined in the outcrops between −1298 m and −1280 m; these shales have TOC values varying
of the southern flank of the Murzuq Basin and to its corresponding between 5% and 13% and are correlated to distinctly elevated Gamma-
subcrops in southwestern Libya (Massa, 1988, Fig. 50). ray values in the wireline log (Figs. 5 and 6), corresponding to the “hot
The very base of the Argiles Principales Formation, in borehole Tt1 shales” definition (cf. Lüning et al., 2000). An interesting feature of the
(−1302 m to −1298 m) is characterized by fine-grained sandstones hot shales in borehole Tt1 is the relatively high carbonate content,
and quartzites with interbedded thin shale layers yielding graptolites which was also noted previously on the basis of the calcimetric well

Fig. 4. Stratigraphic occurrences of chitinozoans and miospores (cryptospores and spores triletes) in borehole Tt1, and previously established graptolite biozonation (Jaeger et al.,
1975).
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 383

Fig. 5. Stratigraphic variation of δ13Corg and correlation with variation in relative γ-ray intensity (arbitrary unit) in borehole Tt1.

log data (Jaeger et al., 1975, Fig. 3). Although they were not directly interval (−1267 m to −1247.9 m) was not available for study from
observed in the core samples analyzed for the present study, the borehole Tt1, but according to previous palaeontological analyses
original lithological log of borehole Tt1 compiled by the drilling (graptolites), it includes at least the vulgaris Zone of late Wenlock age
company SEREPT reports the presence of dolomitic beds around − (Jaeger et al., 1975). The uppermost shaly interval ranging between −
1288 m. According to the graptolite data (Jaeger et al., 1975) but also to 1247.9 m and − 1225.2 m (six samples analyzed), has yielded
new data from nearby boreholes (e.g., Lg3, St1; Storch, personal miospores and chitinozoans which indicate precisely a late Ludlow
comm., September 2007), the hot shale interval in borehole Tt1 as well (Ludfordian) age (Figs. 3 and 4). The Silurian sequence is truncated by
as in the adjacent wells in southern Tunisia contains at least the a major erosional surface at ca. −1224 m, above which sediments of
sedgwikii graptolite Zone (Aeronian; Figs. 3 and 4). Moreover, Late Permian age occur (Jaeger et al., 1975, and unpublished reports by
graptolites of the murchisoni and rigidus zones, indicating a Shein- SEREPT oil company).
woodian (early Wenlock) age occur within and just above the Gamma
ray peak, between ca. −1278 m and −1287 m (Jaeger et al., 1975; Figs. 4 4. Analytical methods
and 5). This clearly indicates that the southern Tunisian hot shales are
middle Llandovery to early Wenlock in age. However, no biostrati- The 27 samples available for study were treated for palynological,
graphic evidence for the Telychian (late Llandovery) has found so far palynofacies, organic carbon isotope, and Rock Eval analyses. Twenty
in borehole Tt1, indicating a possible hiatus or highly condensed grams of sediments from each sample were dissolved in HCl and HF to
sediments corresponding to this stage. The lundgreni graptolite Zone isolate the organic fraction. No oxidation was performed on the
(Homerian; late Wenlock) occurs in the uppermost part of the present organic residue. Organic carbon isotopic analysis was performed on
interval, near −1267 m (Jaeger et al., 1975). 40–100 μg of the unfiltered organic-rich residue with an elemental
Finally, the interval between −1267 m and −1224 m consists of a analyser (Carlo-Erba1110) connected online to a ThermoFinnigan
monotonous succession of dark grey silty shales. A 20 meter-thick Delta Plus mass spectrometer, at the isotope laboratory of the Institute
384 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

Fig. 6. Stratigraphic variations of the TOC content and of the Rock-Eval parameters S2 and HI in core Tt1. Also indicated are the gamma-ray, for comparison with the TOC content, as
well as the interpretation of S2 values in terms of source rock potential.

of Geology and Mineralogy of the University of Erlangen, Germany. For instrument, Oil Show Analyser device (Espitalié et al., 1985a,b, 1986) at
carbon isotope values, the conventional δ-notation in permil, relative the Université Paris 6, France. Other parameters given by the Rock-
to V-PDB (Vienna-PDB) is used. For palynological and palynofacies Eval are the petroleum potential (S2 in mg HC per g of rock), the
study, firstly one slide per sample was mounted using unfiltered hydrogen index (HI in mg HC per g of TOC) which depends on the
organic-rich residue, and subsequently the remainder of the residue origin and state of preservation of the organic matter, and the Tmax
was filtered at 10 μm in order to eliminate the finer fraction of the (°C) which is an indicator of thermal maturity of the organic matter.
amorphous organic matter and facilitate the optical microscopy
analyses. Palynofacies analyses of 10-μm filtered residue is usually 5. Results
recommended in the case of very organic-rich sediments (e.g., Batten,
1996), and this procedure was followed herein. For palynofacies 5.1. Palynology and palynofacies
description and terminology, we followed Combaz (1964), Tyson
(1995), and Batten (1996). Three to nine palynological slides for each All samples yielded palynomorphs in variable abundance and
sample were studied in order to reach a 250 palynomorph-count per diversity: acritarchs and prasinophytes (organic-walled marine
sample needed for standardizing the relative abundance calculations microphytoplankton), chitinozoans (marine microplankton), mios-
of the different palynomorph groups. Additionally, residues from pores (dispersal propagules produced by early land plants) and, more
samples particularly rich in organic-walled microfossils were also rarely, cryptospores were observed (Figs. 3 and 4; Plates I and II).
filtered at 50 μm in order to isolate the larger palynomorphs from the Detailed taxonomic evaluation of the various groups of palynomorphs
fine grained debris, and facilitate the taxonomic and biostratigraphic is not among the scopes of the present paper. Here only the major
study. Relative palynomorph abundance was estimated semi-quanti- features of the assemblages in relation to palynofacies characteristics
tatively on the basis of the relative number of palynomorph species of the sediments, the palynomorph chronostratigraphic significance,
counted in one or more slides in order to reach the 250-specimen and their variation in relative abundance are presented and discussed.
count. Taking into account that all residues were obtained from the The preservation of the various types of organic-walled micro-
same volume of sediment and that slides were mounted with an equal fossils was variable, ranging from very good and light yellow colours of
amount of concentrated organic residue, this procedure is considered vesicle walls, to medium and brownish colours (Plates I–III). No
to be sufficiently reliable to estimate the variations in relative opaque, completely carbonized palynomorphs were observed. An
abundance of palynomoprhs per unity of volume of sediment exception is the preservation of acritarchs and chitinozoans in the
throughout the section (cf. De Vernal et al., 1987). samples corresponding to sediments with the highest TOC values and
The total organic carbon content (TOC in weight %) was highest proportion of amorphous organic matter (AOM) which,
determined on 100 mg of ground sediment sample with a Rock-Eval although presenting only slight thermal alteration, were highly
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 385

Plate I (caption on page 388).


386 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

Plate II (caption on page 388).


M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 387

corroded on their surface, totally masking the microsculptural been also observed, showing provenance from sediments of Middle
elements and the fine branching of small appendices. Changes in Ordovician age, such as species of Frankea and Dicrodiacrodium. These
palynofacies type and relative abundance of the different classes of reworked forms, although generally well preserved, are clearly
sedimentary organic matter (SOM), closely correspond to the main distinguishable from the in situ palynomorphs (apart from their age
changes in palynomorph composition (Figs. 3, 4 and 8). difference with respect to the majority of the assemblage) by their
The palynofacies of samples 1330.0, 1313.6 and 1313.1 (uppermost greater corrosion of the vesicle wall and incomplete/broken fine
Djeffara Formation) are dominated by relatively well preserved, light- appendages (Plate I, figs. 10, 12). The amount of reworked specimens
coloured palynomorphs represented by abundant acritarchs, com- in the assemblages has been estimated to not exceed 2% of the total
monly occurring chitinozoans, and rare cryptospores (Figs. 3, 4; Plates palynomorph content. The presence of reworked palynomorphs has
I and III). Amorphous organic matter (AOM) is subordinate, light in been recorded repeatedly in Hirnantian, glacial-related deposits in
colour, and badly preserved, showing signs of corrosion, probably due North Africa (Paris et al., 2000; Vecoli and Le Hérissé, 2004; Vecoli,
to oxidation (Plate III, fig. 1). 2008). Cryptospores also occur, albeit rarely (ca. 1% in relative
The diverse acritarch assemblages, include the well known species abundance), in this interval, indicating a relatively proximal character
Acanthodiacrodium crassus, Leprotolypa evexa, Neoveryhachium carmi- of the sediments. Chitinozoans are infrequent and badly preserved
nae, Ordovicidium sp., Orthosphaeridium insculptum, O. rectangulare, (mostly broken and with fine structures heavily corroded) so that they
Villosacapsula villosapellicula, clearly indicating a Late Ordovician age cannot be identified at species level with certainty. Specimens of An-
(Fig. 3; Plate I). Various species of large Baltisphaeridium also occur gochitina sp., Belonechitina sp., Conochitina sp., Conochitina cf. elon-
which are consistent with this age attribution. Moreover, the presence gata, and Lagenochitina sp. were recorded. The chitinozoan
of Dactylofusa cucurbita, Deunffia sp., Hoegklintia sp., Saharidia assemblage, although compatible with a Late Ordovician age, does
munfarida strongly supports a latest Ordovician (Hirnantian) age of not permit to further refine the chronostratigraphic attribution
the sediments (Fig. 3; and cf. Vecoli and Le Hérissé, 2004; Vecoli, provided by the acritarchs in this interval.
2008). Acritarchs are generally well preserved, their walls being The palynofacies type and palynomorph assemblages abruptly
transparent and pale yellow in colour. However, some minor signs of change starting from sample 1301.5, at the base of the Argiles
corrosion of the wall are visible which may be attributed to slight Principales Formation (Rhuddanian; early Llandovery; Figs. 3, 4
oxidation within the sediment (Plate I). Reworked acritarchs have and 8). Between −1301.5 m and −1294.0 m the dominant fraction of

Plate III (caption on page 388).


388 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

the palynological residue is AOM, and palynomorphs occur rarely and irregularly distributed on vesicle surface. An acritarch diversity peak is
are much less diverse than in the preceding interval. Acritarchs are observed at −1291.1, where species of Oppilatala, Multiplicisphaeri-
virtually absent and only very rare triangular Veryhachium spp. and dium, and Leiofusa occur. The diversity of acritarch assemblages and
Evittia spp. were observed. These show dark yellow to brown vesicles palynomorph vs. AOM relative abundance slightly increase again in
and corroded appendixes. Prasinophytes are also rare within this samples 1273.0, 1270.5, reaching a further peak at −1267.0 m, were
interval, being represented by, Tasmanites, Leiosphaeridia and Cyma- numerous acritarch taxa first occur (Fig. 3). At this level, the
tiosphaera in decreasing order of abundance (Figs. 3, 4). In general, prasinophyte assemblage is also more diverse than in stratigraphically
abundances of acritarchs and of prasinophycean algae are inversely lower levels, due the inception of species of Dictyotidium, Quadrati-
correlated. The latter are much more abundant than the former tum, and Duvernaysphaera. The precise chronostratigraphic signifi-
between −1297 m and −1282.0 m, the interval that corresponds to the cance of the phytoplankton assemblages recovered from the interval
more organic-rich shales, with the highest values of natural radio- between −1301.5 m and −1267.0 m is difficult to evaluate because the
activity (Figs. 3, 5, 8). An important fraction of the insoluble exact stratigraphic ranges of many species are still poorly known.
palynological residue is constituted by pyrite, often occurring in However, a general agreement with the biostratigraphic ages provided
very fine framboidal aggregates. In all the basal part of the Argiles by the graptolites is evident. In particular, the species Ammonidium
Principales Formation (−1301.5 m up to −1267.0 m) AOM is relatively microcladum, Oppilatala sp., Tylotopalla sp., Quadraditum fantasticum
well preserved (Plate III, fig. 2), dark yellow to brown in colour and are consistent with an early to middle Silurian (Llandovery–Wenlock)
generally poorly fluorescent. It occurs abundantly in generally large age.
but dimensionally variable aggregates of extremely fine particles Chitinozoans occur in generally low but constant abundances
clearly discernible under UV light. Observation in fluorescent light throughout the stratigraphic interval between −1301.5 m and −
also revealed, in some specific levels, a significant number of 1267.0 m, and show variable preservation from highly corroded to
palynomorphs within the groundmass of the more densely aggre- well preserved vesicles. The following stratigraphically significant
gated particles which appeared semi-opaque at their centre upon species have been observed (Fig. 4): Spinachitina fragilis (which is an
optical microscopy observation (Plate III, fig. 4). Neither structure, nor index species for the basal Rhuddanian) occurs at −1301.5 m. Spina-
remnants of possible pre-existing structures are detectable even if chitina maennili and Conochitina edjelensis edjelensis are present
observed in fluorescent light. between − 1296.0 m and − 1288.35 m thus indicating a latest
The interval between − 1291.1 m and − 1267 m is richer in Rhuddanian–middle Aeronian age. Angochitina longicollis, whose
palynomorphs; acritarch assemblages here are better preserved and known stratigraphic range is from Telychian −lower Sheinwoodian,
dominated by species of Ammonidium (e.g., Ammonidium microcla- occurs at −1286.9 m. The above chitinozoan datings are based on the
dum), and, secondarily, of Tylotopalla and Leiofusa. Tylotopalla is often synthetic chitinozoan range chart for the Silurian compiled by Paris
represented by aberrant forms showing few, heteromorphic processes (1996) and on data from the global Silurian chitinozoan zonation

Plate I. Latest Ordovician acritarchs and cryptospore (sample Tt1-1313.1). Scale bar equals 10 μm. (see page 385)

Fig. 1. Multiplicisphaeridium sp.


Fig. 2. Acanthodiacrodium crassus (Loeblich and Tappan, 1978) Vecoli, 1999.
Fig. 3. Veryhachium lairdii (Deflandre) Deunff, 1959 ex Downie, 1959.
Fig. 4. Cryptospore.
Fig. 5. Dactylofusa striatogranulata Jardiné et al. (1974).
Fig. 6. Leprotolypa evexa Colbath, 1979.
Fig. 7. Multiplicisphaeridium sp.
Fig. 8. Deunffia sp.
Fig. 9. Villosacapsula setosapellicula (Loeblich) Loeblich and Tappan, 1976.
Fig. 10. Dicrodiacrodium ancoriformis Burmann, 1968 emend. Servais et al., 1996⁎.
Fig. 11. Quadruilobus spinatus Tappan and Loeblich, 1971.
Fig. 12. Frankea sartbernardensis (Martin) Colbath, 1986⁎.

Plate II. Chitinozoans, miospores and acritarchs of Ludlow (late Silurian) age (samples Tt1-1247.9 and Tt1-1225.2). Scale bar equals 10 μm. (see page 386)

Fig. 1. Angochitina echinata Eisenack.


Fig. 2. Cingulochitina sp. A of Paris, 1981.
Fig. 3. Ancyrochitina primitiva Eisenack.
Fig. 4. ?Archaeozonotriletes sp. cf. A. divellomedium Chibrikova, 1959.
Fig. 5. cf. Synorisporites verrucatus (Richardson and Lister) Richardson and Ioannides, 1973.
Fig. 6. Ambitisporites avitus Hoffmeister, 1959.
Fig. 7. Baltisphaeridium sp.
Fig. 8. Dateriocradus polydactylus Tappan and Loeblich, 1971.
Fig. 9. Neoveryhachium carminae (Cramer) Cramer, 1971.
Fig. 10. Oppilatala sp. cf. O. eoplanktonica (Eisenack) Dorning, 1981.
Fig. 11. Evittia remota (Deunff) Lister, 1970.
Fig. 12. Leiofusa algerensis Cramer, 1970 ex Eisenack et al., 1976.

Plate III. Palynofacies of latest Ordovician and Silurian sediments in borehole Tt1. Scale bar equals 50 μm in Figs. 1–3, and 10 μm in Figs. 4–12. (see page 387)

Fig. 1. Palynofacies of sediments from uppermost Djeffara Formation, latest Ordovician (sample Tt1-1313.1). Note well preserved and abundant acritarchs, subordinate
AOM, light in colour and badly preserved.
Fig. 2. Palynofacies of sediments from the “hot shale” interval in the Argiles Principales Formation, Aeronian, Middle Llandovery (sample Tt1-1291.1). Note extreme
abundance of well preserved, structureless AOM.
Fig. 3. Palynofacies of sediments from the upper part of the Argiles Principales Formation, late Ludlow (sample 1240.5). Note well preserved and diversified palynomorph
assemblages, including acritarchs and miospores, and subordinate AOM.
Fig. 4. Fluorescent light view of structureless, partly fluorescent well preserved AOM from the “hot shales” interval in borehole Tt1 (sample Tt1-1291.1), revealing highly
fluorescent acritarch completely embedded within large AOM aggregate, and indiscernible under normal light microscopy observation.
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 389

proposed by Verniers et al. (1995); they are in remarkable agreement


with the ages inferred from graptolites.
A further marked change in the palynomorph composition and
palynofacies type occurs at −1247.9 m, characterizing the rest of the
stratigraphic section, up to − 1225.2 m just below the major
unconformity surface. In this interval, abundance of palynomorphs
increases again, while the amount of AOM is reduced (Plate III, fig. 3).
One of the most striking features is the abundant occurrence of land-
derived organic debris and palynomorphs (miospores) which repre-
sent ca. 30% of the total palynomorph content (Fig. 8). Several
cryptospores and trilete spores species are present, of which the most
abundant and biostratigraphically significant are Ambitisporites avitus,
?Archaeozonotriletes sp. cf. A. divellomedium, Laevolancis divellome-
dium, Retusotriletes warringtoni, and Synorisporites verrucatus, and
Emphanisporites rotatus. The co-occurrence of these taxa permits to
identify the Synosporites lybicus – Chelinospora poecilomorpha Zone,
spanning the late Gorstian – Ludfordian time interval. Acritarchs are
represented by well preserved and diversified assemblages, including
some long-ranging taxa (Late Ordovician – Silurian) species such as
Evittia remota (sensu “Baltisphaeridium denticulatum complex” of
Cramer, 1970); Neoveryhachium carminae, various species of Balti-
sphaeridium, but also some taxa characteristic of middle-late Silurian
age, such as species of Visbysphaera, Oppilatala, Triangulina, Tuni-
sphaeridium. Other taxa have more restricted stratigraphic ranges,
Fig. 7. Plot of HI vs Tmax values for core Tt1 (diagram from Espitalié et al., 1986). Symbols
such as Deflandrastrum millepiedi and D. leonardi, which permit to
as in Fig. 6.
tentatively restrict the age range of the assemblage to the late Ludlow
(Ludfordian; e.g., Cramer, 1970), further refining the biostratigraphic
age provided by the miospore assemblage.
Well preserved and relatively diversified chitinozoans are also excursions evidencing three maxima at −1294 m (+2‰ positive shift;
present, including the chronostratigraphically significant species An- hot shale interval; early Aeronian, middle Llandovery), at −1286.9 m
cyrochitina primitiva, Angochitina echinata (Plate II, Figs. 1, 2), Cingu- (+2.5‰ positive shift; hot shale interval; Llandovery/Wenlock bound-
lochitina cf. convexa, and Sphaerochitina acanthifera, well compatible ary or earliest Wenlock), and at −1240.5 m (+3‰ positive shift; shaly
with the late Ludlow age suggested by acritarch evidence (Verniers level within the predominantly silty upper portion of the Argiles
et al., 1995; Paris, 1996). Principales Formation, Ludfordian, late Ludlow). The minimum values
are mostly concentrated in the lower part of the section, correspond-
5.2. Toc and rock-eval ing to the poorly sorted sandstones and silts of the Djeffara Formation
(latest Ordovician) and to the fine grained sandstones at the base of
The TOC values range between 0.2 and 13.2% (Fig. 6). The vertical the Argiles Principales Formation (earliest Llandovery). In addition,
distribution of TOC values is clearly correlated with the Gamma-ray very low δ13Corg are also recorded in a short interval just above the hot
curve (Fig. 6). Very low TOC values (average 0.4%) are observed in the shale package within the Argiles Principales Fm. (Wenlock). A much
Hirnantian sediments of the Djeffara Formation, between −1330 m less pronounced and more gradual excursion (+1‰ positive shift;
and −1313.1 m. An increase is noted at the transition with the Argiles maximum δ13Corg value −29.83‰) is also observed at −1273 m
Principales formation which is generally characterised by high TOC (possibly middle Wenlock; Sheinwoodian/Homerian transition). An
values (5.6%). Maximum TOC values identify the hot shale interval isolated, relatively high δ13Corg value (−28.87‰) finally occurs at the
(−1298 m to −1280 m). A return to relatively low TOC values is topmost level of the study section, only a few tens of centimetres
observed in the Ludlow silty shales (average 1.4%), except at − below the conspicuous discontinuity surface at −1224 m.
1240.5 m, where TOC values up to 2.9% are recorded (Fig. 6).
Tmax values range between 428 and 443 °C (Fig. 7), with an average 6. Discussion
of 437 °C. These values indicate that the organic matter has just
reached the beginning of the oil window. The present data evidence remarkable correlation between
The type-and therefore the origin- of the organic matter present palaeontological, palynofacies, isotopic, and lithological develop-
can be determined using the HI-Tmax diagram (Fig. 7; Espitalié et al., ments throughout the study section (Figs. 3–8). Palynofacies of the
1986). Within this diagram, most of the samples plot in the lower part Hirnantian sediments of the Djeffara Fm., showing low abundances of
of Type II organic matter while a few samples plot in the Type III part highly degraded AOM, suggest a well oxygenated environment; this is
of the diagram. also well supported by organic geochemistry data (low TOC and HI
Consistent with the high TOC values and relatively low thermal values). Moreover, the high diversity of the acritarch assemblage
maturity of the samples, a high petroleum potential (S2) is observed in indicates normal marine, oligotrophic conditions, relatable to outer
most of the samples of the Argiles Principales Formation which rates shelf palaeoenvironmental setting. In the literature, the Hirnantian
as good to very good petroleum source rock. Conversely, the sediments of the “Argiles microconglomératiques”-type are often
Hirnantian Djeffara Formation and the Ludlow silty shales only considered of peri-glacial nature, deposited mostly during the melting
show poor to fair source rock potential. of the Gondwanan ice-cap (Paris et al., 1995, 2000), which is
consistent with well oxygenated and normal marine environment of
5.3. Stable carbon isotopes deposition. The low isotopic values recorded in the three samples
analyzed in this interval (Fig. 5) possibly indicate a late Hirnantian
In the Tt1 core (Fig. 5), a baseline of low δ13Corg values varying time (i.e., post- early Hirnantian isotopic excursion), and thus are also
between −30.79 and −30.59‰ is interrupted by three distinct positive consistent with this scenario.
390 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

Fig. 8. Stratigraphic variations of palynofacies type, relative abundance of main classes of palynomorphs, and diversity of marine microphytoplankton.

There is no evidence of major discontinuities at the Ordovician/ only partial preservation of OM. Alternatively, anoxia was effectively
Silurian transition in the study section, even if the sequence is very developed within the sediments, but the sea level was shallow enough
condensed (Figs. 3 and 4); this is clearly substantiated by the presence to keep the sediment/water interface above the average storm wave
of Rhuddanian (earliest Silurian) graptolites (vesiculosus Zone) base, which would have caused periodical remixing and oxygenation
between −1300.4 m and −1301.7 m and chitinozoans (fragilis Zone) of the sediment. This possibility seems supported by the presence of
at −1300 m, a few meters above the last occurrence of Hirnantian silty to sandy cm-thick layers within a predominantly shaly succes-
(latest Ordovician) palynomorphs (Figs. 3 and 4). The distinct sion, but also by the low values of the HI index of the OM (Figs. 3
lithological change marking the boundary between the Djeffara and and 7). It should be borne in mind that this basal interval of the Argiles
Argiles Principales formations (ca. −1302 m; Fig. 3) coincides with a Principales Fm. is Rhuddanian in age, meaning that is was being
first important change in palynofacies type and turnover of the deposited while in the deeper parts of the Ghadamis Basin (but also in
palynomorph assemblage (Fig. 8). The basal portion of the Argiles other basins in adjoining areas; Fig. 1) extensive black shale deposition
Principales Formation (−1302 m to −1298 m) is characterized by was already active and thus anoxic conditions were widespread across
palynofacies dominated by slightly degraded and essentially non- the entire Sahara Platform (Lüning et al., 2000).
fluorescent AOM, by a drastic reduction in diversity of the acritarch The base of the hot shale interval, at −1298 m, marks a second well
assemblage and by an increase in occurrence of prasinophyceae (Figs. defined change in palynofacies, petrological, and organic geochemical
3 and 8). TOC values increase markedly, but values of HI as well as δ13C parameters (Figs. 3–8). In the palynofacies spectrum, a further
values remains low and similar to the values recorded in the increase in abundance of AOM is recorded. Hot shale AOM still
Hirnantian sediments of the Djeffara Fm. (Figs. 5–8). Globally, these appears as featureless aggregates of very fine particles (≤1 μm in
characteristics suggest an increase in productivity, although palaeoen- diameter), but shows more elevated fluorescence, suggesting
vironmental conditions were not particularly favourable to preserva- enhanced preservation of organic matter (cf. Tyson, 1995). Increased
tion of organic matter. Two possible scenarios can be envisaged to values of HI and exceptionally high TOC values (consistent with the
explain this situation. A first possibility is that the increase in increased fluorescence of the AOM; Tyson, 1995, 2006) coupled to
bioproductivity was still not strong enough to cause complete anoxia elevated levels of natural radioactivity (Figs. 6 and 7) also indicate
within the sediment which only reached a dysoxic state, permitting better preservation of AOM, thus suggesting the establishment of full
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 391

anoxic conditions within the sediments. The low diversity of the δ13Corg data presented by Cramer and Saltzman (2007b, Fig. 4) do not
acritarch assemblages, associated to an increase in the abundance of depict an evident negative excursion, but rather seem to be scattered
prasinophyceae algae, strongly support enhanced surface bioproduc- over a wide range of values. More isotopic data on organic carbon will
tivity and establishment of eutrophic conditions in surface oceanic be needed in order to verify and correctly interpret the apparently
waters (Combaz, 1966; Revill et al., 1994). According to Lüning et al. divergent isotopic signatures in the different species of carbon as
(2000), conditions of strong upwelling all along the margin of presented by Cramer and Saltzman (2007b). For the scope of the
Gondwana at high latitudes were at the origin of extensive organic- present discussion, we will therefore assume that our δ13Corg
rich black shales deposits in large parts of the Sahara Platform, the excursion effectively records the high-latitude expression of the well
middle East and in the Arabian Peninsula. The present results are in known excursion which correlates with the Ireviken Event. Therefore,
agreement with a scenario of upwelling-triggered high surface our data can be discussed in the light of the oceanographic and
productivity and anoxia in the water column and in the bottom climatic models proposed to account for the observed changes in the
sediments for the origin of the Tunisian hot shales. It should also be Silurian isotopic, sedimentary, and fossil records (Jeppsson, 1987,
noted, that the abundance of prasinophyceae may indicate prolifera- 1990, 1993, 1997; Munnecke et al., 2003). Recently, Cramer and
tion in upwelling nutrient-enriched waters (Revill et al., 1994). The Saltzman (2007a), presented a review of these models, with a new
observed offset in hot shale depositional age (younger on palaeo- proposal which has been developed in order to explain the apparent
highs, older in basin palaeo-depressions) might be well explained by contradiction between increased levels of organic carbon burial
deposition of organic-rich sediments during a marine transgression, required to explain the positive isotopic shift of the Ireviken Event,
as postulated also by Lüning et al. (2000). Such temporal changes in and the widespread development of carbonate platforms in low
the location of upwelling-derived organic-rich sediments in parallel to latitudes during early to middlle Silurian times. These authors
sea level variations have been well documented in recent situations suggested that the Ireviken Event developed during a time of
(Mollenhauer et al., 2002). The hot shale interval is also characterized decreasing levels of primary productivity in epeiric seas due to a
by two distinct isotopic excursions, one occurring in levels dated as large transgression linked to the demise of a late Llandovery
Aeronian (middle Llandovery) and the other, more pronounced, glaciation. At the transgression peak, a climatic warming would
occurring in levels dated as Sheinwoodian (basal Wenlock; Fig. 5). have caused not only a shift from estuarine to anti-estuarine
Early to middle Silurian isotopic excursions are well known in circulation in epeiric seas favouring the development of carbonatic
carbonate sequences from many low palaeolatitude settings (Laur- factories, but also a change in the site of deep water formation from
entia and Baltica), where they are often associated to faunal high- to low-latitude setting, with consequent development of anoxia
extinctions and lithological changes (see Munnecke et al., 2003 and in the oceanic deeper basins. The decoupling between carbonatic
Cramer and Saltzman, 2007a, for a review). One of the most studied of epicontinental platforms and the deep anoxic oceans (sites of
these isotopic events is the pronounced early Sheinwoodian (earliest sequestration of isotopically light carbon) would largely explain why
Wenlock) excursion, associated with the so-called “Ireviken Event” the early-middle Silurian isotopic excursions are found to occur during
(Jeppsson, 1987, 1990, 1993, 1997; Munnecke et al., 2003; Cramer and intervals of increased carbonate production.
Saltzman, 2005; Kaljo and Martma, 2006; Cramer and Saltzman, In the Tt1 section, palaeogeographically located on a basin-margin
2007a,b; Loydell, 2007). This event appears to be global in character, palaeohigh, the main isotopic peak (at −1286.9 m) occurs within the
as it has been recorded in several areas (Fig. 2), especially in equatorial hot shale interval, near its upper end. It is then followed by a return to
to tropical latitude settings such as North America (Saltzman, 2001; organically leaner silts and shales, probably meaning that the bulk of
Cramer and Saltzman, 2005; Cramer et al., 2006; Cramer and organic matter burial actually took place in high latitude settings
Saltzman, 2007b), the Baltic region (Kaljo et al., 1997; Munnecke slightly before the onset of the “Ireviken” isotopic excursion. This is
et al., 2003; Martma et al., 2005), and Australia (Talent et al., 1993). especially clear considering that in the deeper parts of the Gondwanan
Although Silurian carbonate deposits are not extensively present in intracratonic basins, hot shale deposition occurred significantly earlier
the Ghadamis Basin, major global palaeoenvironmental disturbances than on basin margins (in Rhuddanian, early Llandovery times). It
should be detectable in the isotopic signature of the sedimentary must be noted also that the occurrence of thin dolomitic layers
organic carbon. Previous studies have in fact clearly demonstrated interbedded within the Tunisian hot shales such as around −1288 m in
that if variations in δ13C are the expression of changes in carbon borehole Tt1, may be a consequence of the large availability of Ca and
cycling at a global scale, not only are they detectable in both inorganic Mg during this time interval. The apparent absence of Telychian
and organic carbon species, but also show the same variation polarity sediments in the Tt1 section (Figs. 3 and 4) might be in agreement
(Strauss et al., 1997; Hayes et al., 1999). Principally because of its with a climatically (glacially?) driven sea level lowstand in late
stratigraphic position, precisely established in early Sheinwoodian Llandovery times, as postulated by Cramer and Saltzman (2007a), but
times, the more pronounced and stratigraphically highest of the two obvious direct evidence of a middle Silurian glaciation is lacking in our
isotopic excursions observed in the hot shale interval in borehole Tt1, section and generally in North Africa. Additionally, the present results
might represent the record of the “Ireviken Event” in the study area. prove that important volumes of highly organic-rich sediments were
An isolated peak in acritarch diversity is recorded in one sampling effectively deposited in Gondwanan high latitude intracratonic
level about 4 m below the isotope peak, but the significance of this shallow seas during an extended period of time from basal Llandovery
diversity fluctuation is difficult to evaluate because of the generally to basal Wenlock. This evidence partly contradicts Cramer and
very low abundance of acritarchs in these levels. In the Baltic regions, Saltzman (2007a) assertion that the Ireviken excursion is associated
the Ireviken Event is usually preceded by at least two minor isotopic with a “decrease in organic carbon burial in shallow water globally”.
excursions in δ13Ccarb, of Aeronian and Telychian time, respectively A progressive return to lower productivity and less anoxic
(Munnecke et al., 2003). Consistently, the early Sheinwoodian conditions after the end of hot shale deposition is suggested in
excursion is preceded in the Tt1 section, by a smaller excursion borehole Tt1 by increasing diversity of the microphytoplankton
dated as middle Llandovery (Aeronian), while the absence of a assemblages accompanied by decreasing abundance of poorly
Telychian excursion may be well explained in our case by the hiatus fluorescent AOM as well as decreasing TOC (and δ13Corg values) in
embracing the major part, if not the entirety of the upper Llandovery the interval between −1278.6 m and −1267.0 m (Figs. 3 and 7). This
(Figs. 3–5). These results are apparently in contrast with the recent stratigraphic interval shows similar characteristics, and might some-
study by Cramer and Saltzman (2007b) who described a trend how reflect comparable palaeoenvironmental conditions of the basal
towards more negative δ13Corg values through the Ireviken event in part of the formation, between −1302 m and −1298 m, though still
the North American Midcontinent. However, in our opinion, the generally high HI values testify to a relatively good preservation of OM.
392 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

This decrease in surface productivity could parallel the documented water column are very difficult to evaluate, because: 1) it is possible
3rd (2nd?) order sea-level fall occurring in the upper Wenlock (Lüning that microphytoplankton did not produced cysts during specific
et al., 2000). environmental conditions; 2) the absolute abundance of cysts per unit
The upper part of the section is characterized by a large positive of volume of sediments is strongly dependent by hydrodynamic
excursion in δ13Corg, which coincides with a slight increase in TOC factors prevailing during deposition; 3) possibility of selective post-
within an interval of generally low TOC and low HI values (−1247.9 m diagenetic degradation/preservation of organic cysts within the
to −1225.2 m; Figs. 5 and 7). The associated palynofacies shows sediment. Tongiorgi et al. (2003), for example, showed that the
abundance of palynomorphs (miospores, prasinophytes, acritarchs number of acritarchs per gram of rock in Middle Ordovician strata of
and chitinozoans) and an essentially non-fluorescent AOM. The the Yangtze Platform, China, was strictly controlled by lithology, and
acritarchs are in general relatively abundant and diverse (Fig. 3), that no definite meaning could be attached to the distribution of
suggesting conditions of normal marine platformal environment palynologically barren samples. The present results show, however, an
throughout this interval. A marked decrease in acritarch diversity is evident similarity with the observations made by Stricanne et al.
recorded starting at −1243.0 m, immediately before the carbon (2006) on compositional changes of acritarch assemblages across the
isotopic maximum (Figs. 5 and 8). A distinct change in the Lau event at Gotland, Sweden in that a major acritarch turnover is
composition of the acritarch assemblages is associated with this recorded to occur during the increase in isotopic values (Figs. 3, 5
diversity decline (Figs. 3 and 8). Post-excursion acritarch assemblages and 8). However, the dominance of acritarchs with short, non–
are not only less diverse, but are largely dominated by a few species ramified processes in correspondence of the late Ludfordian excursion
belonging to the genera Evittia, Neoveryhachium, and by triangular which characterizes the Swedish acritarch assemblages at Gotland
Veryhachium, with the other taxa represented by very rare occur- (Stricanne et al., 2006) was not observed in the present study. This
rences. Miospores show an evident increase in abundance in parallel might reflect different palaeobiogeographic settings (low versus high
with the increasing values of δ13Corg, reaching a maximum in palaeolatitudes) of the two areas.
correspondence of the isotope peak, where they make up ca. 30% of The recorded input of miospores in the Tt1 borehole reflects two
the total palynomorph content. The geochemical as well as palyno- concurrent factors: one is the evolution of early land plants, with a
facies parameters can be interpreted as to indicate oligotrophic progressive increase in geographical diffusion, biological complexity
conditions of the surface waters and well oxygenated bottom and diversity of spore-producing plants during the critical Silurian
sediments. The marked isotopic excursion at −1240.5 m falls in a time interval. The second factor is the progradation of shallow and
stratigraphic interval which is dated as late Ludlow based on acritarch, marginal marine deposits into more basinal areas during a second-
chitinozoan and miospore evidence (Figs. 3–5). A major late Silurian order regression observed regionally over the entire North Africa (HST
(late Ludlow, Ludfordian) isotopic excursion is well known from many of sequence NA2 of Carr, 2002). At the present state of knowledge, it
localities of low palaeo-latitudes (Fig. 2) in Baltica (Sweden and Baltic would be premature to speculate about palaeoclimatic variations
States: Samtleben et al., 1996, 2000, Martma et al., 2005), Laurentia (humid vs. arid conditions) based on the spore assemblages. The
(North America: Saltzman, 2001), and eastern Gondwana (Australia; occurrence of relatively abundant miospore content in the palynolo-
Andrew et al., 1994; Jeppsson et al., 2007), as well as from one middle gical residues is not likely to have a significant influence on the carbon
palaeo-latitude peri-Gondwana locality of the Prague Basin (Lehnert isotopic signal, as it might be inferred from the apparent correlation
et al., 2007). This prominent excursion is correlated to lithological and between the δ13C Ludfordian peak and the maximum in miospore
faunal changes in many sections worldwide and it is generally abundance at 1240.5 m (Figs. 4 and 8). Differences in the δ13C between
associated to an oceanic event named the “Lau Event” (see recent Silurian marine palynomorphs (acritarchs and prasinophyte algae)
review in Martma et al., 2005). Although the precise age of the and terrestrially derived woody fragments of Devonian age were
excursion is currently debated between a middle or late Ludfordian documented by Lécuyer and Paris (1997). These authors measured
(late Ludlow) age (Kaljo and Martma, 2006; Lehnert et al., 2007), there significant differences, up to 4‰, between the carbon isotopic ratio of
is little doubt that the large isotopic excursion occurring in the upper land plant derived woody fragments and marine algae, although these
part of the Tt1 section (broadly dated as late Ludlow) is correlatable to were measured in specimens extracted from samples of different age
the Lau Event. and localities. In our material, at the maximum abundance, miospores
As with the other Silurian excursions, the Lau Event has been contribute for the 30% of the palynological content, 60% of it being of
related to changes in palaeoclimatic conditions (Martma et al., 2005; marine derivation (phytoplankton). In addition, the bulk of the
Lehnert et al., 2007), switching from a humid to an arid state. organic residue (60%) is constituted of amorphous organic matter of
Interestingly, a recent detailed study which included analyses of marine origin (derived either from bacterioplankton or degraded
palynological changes across the Lau event in Sweden (Stricanne et al., phytoplankton). Lécuyer and Paris (1997) calculated that 25% of
2006) recorded a peak in abundance of miospores in correspondence woody tissue in an organic residue would modify the δ13C value of the
of the carbon isotope peak, similarly to what observed in borehole Tt1. sediment by 1‰. Accordingly, a significant effect on the carbon
In the latter study, acritarch abundance was anti-correlated with isotopic fractionation curve would be expected only if woody
miospore abundance and this was taken as an indication of decreasing fragments occurred abundantly in the palynological residue, as it
levels of primary production in the photic zone. The input from has been effectively observed in sediments of Devonian age
terrestrial plant sources (the miospores) was explained by Stricanne (Joachimsky, pers. comm., February 2008). However, in our study
et al. (2006) as due to aeolian transport because other possible samples, woody fragments are absent, and only miospores occur in
explanations (sea-level drop or instauration of humid climate) were maximum abundance of ca. 15% of the total organic residue. Miospores
not supported by sedimentological and isotopic (e.g., δ18O) evidence. are mostly very thin-walled (1–2 μm), spherical vesicles which are
On the basis of the available data, we are not able either to confirm or empty internally, and thus contribute much less towards the total
reject this hypothesis, because lithological and sedimentological mass of atoms of C per unit of volume than woody fragments.
indicators of climatic conditions are not obvious in the predominantly The difficulty of explaining the driving mechanisms of the Lau
fine-clastic sedimentary sequence studied by us (not continuously Event and associated isotopic excursion with the same oceanographic
cored). However, integrating the geochemical and palynofacies models proposed for the early to middle Silurian events (e.g., the
evidence, it is possible to infer “normal” conditions of palaeoproduc- Ireviken event) was recently highlighted by a new detailed study of
tivity as well as oxygenated bottom sediments, moderately favourable late Ludfordian sediments from Gotland, Sweden (Eriksson and
to preservation of organic matter. The relationships between acritarch Calner, 2008). These authors proposed two possible mechanisms for
abundance in the sediments and levels of primary production in the the late Ludfordian isotope excursion: 1 — a glacio-eustatically
M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394 393

induced increase in weathering rate due to lowered sea level causing a 5. The Lau event is more difficult to interpret. Extensive black shale
change in the riverine C-weathering flux towards enriched δ13C values deposition does not occur in association with this event, and levels
("weathering hypothesis" of Kump et al., 1999); and 2 — an increased of primary productivity seem to have remained essentially
photosynthetic activity by benthic cyanobacteria causing enrichment unchanged before and during the event (oligotrophic conditions
in δ13C in precipitating carbonates. It is interesting to note that the as suggested by palynofacies evidence). An input of organic matter
miospore abundance peak we observe at the Lau event would be a of terrestrial derivation into the marine sedimentary record
consequence of the former scenario and not of the latter. (miospores of early land plants) correlate exactly with the
The isotopic excursions recorded here have generally smaller maximum values of organic carbon isotopic fractionation, similarly
amplitudes than those corresponding to the same events but to what has been observed in the Baltic region (Stricanne et al.,
measured in carbonatic sections elsewhere. This is particularly 2006).
evident for the Ludfordian (Lau) excursion, which in the literature is 6. Our study demonstrate that previous models developed in the low
described as having an average amplitude between 8–9‰, exception- to intermediate palaeolatitude settings to explain the observed
ally reaching 10‰ δ13C. In comparison, our reported Ludfordian association among isotopic development, facies changes, and
excursion of about 3‰ δ13C is significantly smaller. A trend towards a fluctuation in biodiversity, cannot be readily applied in the more
basinward decline in the magnitudes of positive δ13C excursions has general case, and particularly in the case of high-palaeolatitude
been previously described in the literature (e.g., Kaljo et al., 1997; localities. More detailed investigation from different Gondwanan
Noble et al., 2005; Melchin and Holmden, 2006), and was discussed localities will be needed in order to enlarge the dataset and
more in detail by Loydell (2007) for the early and middle Silurian improve our knowledge of the Silurian global palaeoceanographic
excursions. The early Sheinwoodian (Ireviken) excursion in borehole events.
Tt1 has similar magnitude as, for example, the one recorded by Kaljo
et al. (1997) in relatively deep water facies in Latvian localities. The Acknowledgements
graptolite-rich, fine-grained early Wenlock sediments of borehole Tt1
are compatible with a distal platformal depositional setting, as The study material, including original borehole logs, was provided
previously noted by Jaeger et al. (1975), and this could easily explain by Dominique Massa. This study is a contribution to the project
the slightly smaller amplitude of the recorded early Sheinwoodian “ECLIPSE II: The Terrestrialization process: modelling complex inter-
excursion. A similar argument can be advanced to explain the reduced actions at the geosphere–biosphere interface”; financial support by
magnitude of the Ludfordian excursion in borehole Tt1. Even if the the CNRS and INSU is gratefully acknowledged. François Baudin and
excursion falls during a time of a prevailing, second order, regressive Florence Savignac (Université P. et M. Curie, Paris 6) are thanked for
trend, the late Ludlow depositional setting of section Tt1is still a performing the Rock-Eval analyses; Petr Storch (Institute of Geology,
relatively distal one (Jaeger et al., 1975). Possibly, an additional factor Prague, Czech Republic) kindly provided unpublished graptolite data
might be that as a result of the discontinuous sampling across the and his help is gratefully acknowledged. Florentin Paris (CNRS Rennes,
excursion interval, the exact stratigraphic level corresponding to the France) and an anonymous reviewer provided critical comments
peak of the curve may have been missed. useful for the improving of the paper.

7. Conclusions
References

1. In the present study, we have analyzed the relationships between Andrew, A.S., Hamilton, P.J., Mawson, R., Talent, J.A., Whitford, D.J., 1994. Isotopic
lithological changes, fluctuations in biodiversity of marine micro- correlation tools in the mid-Palaeozoic and their relation to extinction events.
Australian Petroleum Exploration Association (APEA) Journal 34, 268–277.
phytoplankton, changes in oceanic productivity, development of Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A.,
anoxia, and organic carbon isotopic excursions in latest Ordovician Abu Salah, A., 2005. Origin, sequence stratigraphy and depositional environment of
through Silurian clastic sequences of the high-latitude regions of an Upper Ordovician (Hirnantian) deglacial black shale, Jordan. Palaeogeography,
Palaeoclimatology, Palaeoecology 220, 273–289.
Gondwana. Armstrong, H.A., Turner, B.R., Makhlouf, I.M., Weedon, G.P., Williams, M., Al Smadi, A.,
2. The two major early Sheinwoodian (basal Wenlock) and Ludfordian Abu Salah, A., 2006. Reply to "Origin, sequence stratigraphy and depositional
(late Ludlow) carbon isotope positive excursions, which are thought to environment of an Upper Ordovician (Hirnantian) deglacial black shale, Jordan".
Palaeogeography, Palaeoclimatology, Palaeoecology 230, 356–360.
reflect global palaeoceanographic changes and are associated to biotic
Balducchi, A., Pommier, G., 1970. Cambrian oil field of Hassi Messaoud, Algeria. Geology
events (Ireviken and Lau event, respectively), have been identified for of giant petroleum fields. Memoir, vol. 14. American Association of Petroleum
the first time in Gondwanan high-latitude settings (North Africa). Geologists, pp. 477–488.
Batten, D.J., 1996. Chapter 26A. Palynofacies and palaeoenvironmental interpretation. In:
3. Our data demonstrate that an extended period of black shale
Jansonius, J., McGregor, D.C. (Eds.), Palynology: principles and applications, vol. 3.
deposition occurred regionally (even if discontinuously in geo- American Association of Stratigraphic Palynologists Foundation, pp. 1011–1064.
graphic extent) over the entire North African margin from Beuf, S., Biju-Duval, B., de Charpal, O., Rognon, P., Gariel, O., Bennacef, A., 1971. Les Grès
Rhuddanian to early Wenlock times. The palynofacies and organic du Paléozoïque Inférieur au Sahara. Institut Français du Pétrole, collection science et
technique du pétrole, Technip editions, 18. Paris.
geochemical evidence support an origin by coastal upwelling- Bourahrouh, A., Paris, F., Elaouad-Debbaj, Z., 2004. Biostratigraphy, biodiversity and
promoted productivity increase. The landward shift in location of palaeoenvironments of the chitinozoans and associated palynomorphs from the
coastal upwelling following a long-term (second order?) transgres- Upper Ordovician of the Central Anti-Atlas, Morocco. Review of Palaeobotany and
Palynology 130, 17–40.
sion might explain the observed pattern of black shale deposition Carr, I.D., 2002. Second-order sequence stratigraphy of the Palaeozoic of North Africa.
and is also well in accordance with a previous model of early Journal of Petroleum Geology 25, 259–280.
Silurian black shale deposition on the North African craton, Combaz, A., 1964. Les palynofacies. Revue de Micropaléontologie 7, 205–218.
Combaz, A., 1966. Remarques sur les niveaux à tasmanacées du Paléozoïque saharien.
previously proposed by Lüning et al. (2000). The Palaeobotanist 15, 29–34.
4. Accordingly, the isotopic shift associated with the Ireviken event Cramer, F.H., 1970. Distribution of selected Silurian acritarchs. An account of the
occurred during the later phases of a protracted period of massive palynostratigraphy and palaeogeography of selected Silurian acritarch taxa. Revista
Española de Micropaleontologia, Número extraordinario, vol. 1. 203 pp.
black shale deposition, and thus of organic carbon burial, on
Cramer, B.D., Saltzman, M.R., 2005. Sequestration of 12C in the deep ocean during the
continental platforms located in high-latitude settings. This could early Wenlock (Silurian) positive carbon isotope excursion. Palaeogeography,
well explain the apparent contradiction between carbonate deposi- Palaeoclimatology, Palaeoecology 219, 333–349.
Cramer, B.D., Saltzman, M.R., 2007a. Fluctuations in epeiric sea carbonate production
tion and carbon isotopic shift towards lighter values widely observed
during Silurian positive carbon isotope excursions: a review of proposed
in low latitude areas (North American Midcontinent, Cramer and palaeoceanographic models. Palaeogeography, Palaeoclimatology, Palaeoecology
Saltzman, 2005, 2007a,b; Baltic regions, Munnecke et al., 2003). 245, 37–45.
394 M. Vecoli et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 378–394

Cramer, B.D., Saltzman, M.R., 2007b. Early Silurian paired δ13Ccarb and δ13Corg analyses Martma, T., Brazauskas, A., Kaljo, D., Kaminskas, D., Musteikis, P., 2005. The Wenlock-
from the Midcontinent of North America: implications for paleoceanography and Ludlow carbon isotope trend in the Vidukle core, Lithuania, and its relations with
paleoclimate. Palaeogeography, Palaeoclimatology, Palaeoecology 256, 195–203. oceanic events. Geological Quarterly 49, 223–234.
Cramer, B.D., Saltzman, M.R., Kleffner, M.A., 2006. Spatial and temporal variability in Massa, D., 1988. Paléozoïque de Libye Occidentale. Stratigraphie et Paléogéographie.
organic carbon burial during global positive δ13Ccarb excursions: new insight from Thèse Sc. Univ. Nice, 1-2, 514 pp., Nice.
high-resolution δ13Ccarb stratigraphy from the type area of the Niagarian (Silurian) Massa, D., Jaeger, H., 1971. Données stratigraphiques sur le Silurien de l'Ouest de la
Provincial Series. Stratigraphy 2 (4), 327–340. Libye. Colloque Ordovicien-Silurien, Brest. Mémoires du Bureau des Recherches
De Vernal, A., Larouche, A., Richard, P.J.H., 1987. Evaluation of palynomorph Géologiques et Minéralogiques 73, 313–321.
concentrations: Do the aliquot and the marker-grain methods yield comparable Melchin, M.J., Holmden, C., 2006. Carbon isotope chemostratigraphy of the Llandovery in
results? Pollen et Spores 29, 291–303. Arctic Canada: implications for global correlation and sea-level change. GFF 128,173–180.
El Arnauti, A., Owens, B., Thusu, B. (Eds.), 1988. Subsurface palynostratigraphy of Mollenhauer, G., Schneider, R.R., Müller, P.J., Spieß, V., Wefer, G., 2002. Glacial/
northeast Libya. Garyounis University Publications, Benghazi – Libya. 275 pp. interglacial variablity in the Benguela upwelling system: Spatial distribution and
Eriksson, M.J., Calner, M., 2008. A sequence stratigraphical model for the Late budgets of organic carbon accumulation. Global Biogeochemical Cycles 16, 1134.
Ludfordian (Silurian) of Gotland, Sweden: implications for timing between changes Munnecke, A., Samtleben, C., Bickert, T., 2003. The Ireviken Event in the lower Silurian of
in sea level, palaeoecology, and the global carbon cycle. Facies 54, 253–276. Gotland, Sweden — relation to similar Palaeozoic and Proterozoic events.
doi:10.1007/s10347-007-0128-y. Palaeogeography, Palaeoclimatology, Palaeoecology 195, 99–124.
Espitalié, J., Deroo, G., Marquis, F., 1985a. La pyrolyse Rock-Eval et ses applications. Noble, P.J., Zimmermann, M.K., Holmden, C., Lenz, A.C., 2005. Early Silurian (Wenlock-
Revue de l'Institut Français du Pétrole 40, 563–579. ian) δ13C profiles from the Cape Phillips Formation, Arctic Canada and their relation
Espitalié, J., Deroo, G., Marquis, F., 1985b. La pyrolyse Rock-Eval et ses applications. to biotic events. Canadian Journal of Earth Sciences 42, 1419–1430.
Revue de l'Institut Français du Pétrole 40, 755–784. Paris, F., 1990. The Ordovician chitinozoan biozones of the Northern Gondwana Domain.
Espitalié, J., Deroo, G., Marquis, F., 1986. La pyrolyse Rock-Eval et ses applications. Revue Review of Palaeobotany and Palynology 66, 181–209.
de l'Institut Français du Pétrole 41, 73–89. Paris, F., 1996. Chapter 17. Chitinozoan biostratigraphy and palaeoecology. In: Jansonius, J.,
Ghienne, J.-F., Boumendjel, K., Paris, F., Videt, B., Racheboeuf, P., Salem, H.A., 2007. The McGregor, D.C. (Eds.), Palynology: principles and applications, vol. 2. American
Cambrian–Ordovician succession in the Ougarta Range (western Algeria, North Africa) Association of Stratigraphic Palynologists Foundation, pp. 531–552.
and interference of the Late Ordovician glaciation on the development of the Lower Paris, F., Elaouad-Debbaj, Z., Jaglin, J.C., Massa, D., Oulebsir, L., 1995. Chitinozoans and Late
Palaeozoic transgression on northern Gondwana. Bulletin of Geosciences 82, 183–214. Ordovician glacial events on Gondwana. In: Cooper, J.D., Droser, M.L., Finney, S.C. (Eds.),
Hayes, J.M., Strauss, H., Kaufman, A.J., 1999. The abundance of 13C in marine organic Ordovician odyssey: short papers for the Seventh International Symposium on the
matter and isotopic fractionation in the global biogeochemical cycle of carbon Ordovician System. The Pacific Section for Sedimentary Geology (SEPM), Fullerton,
during the past 800 Ma. Chemical Geology 161, 103–125. California, pp. 171–176.
Jaeger, H., Bonnefous, J., Massa, D., 1975. Le Silurien en Tunisie; ses relations avec le Silurien Paris, F., Bourahrouh, A., Le Hérissé, A., 2000. The effects of the final stages of the Late
de Libye nord-occidentale. Bulletin de la Société Géologique de France 7, 68–76. Ordovician glaciation on marine palynomorphs (chitinozoans, acritarchs, leio-
Jardiné, S., Combaz, A., Magloire, L., Peniguel, G., Vachey, G., 1974. Distribution spheres) in Well Nl.2 (NE Algerian Sahara). Review of Palaeobotany and Palynology
stratigraphique des acritarches dans le Paléozoïque du Sahara algérien. Review of 113, 84–104.
Palaeobotany and Palynology 18, 99–129. Revill, A.T., Volkman, J.K., O'Leary, T., Summons, R.E., Boreham, C.J., Banks, M.R., Denwer,
Jeppsson, L., 1987. Lithological and conodont distributional evidence for episodes of K., 1994. Hydrocarbon biomarkers, thermal maturity, and depositional setting of
anomalous oceanic conditions during the Silurian. In: Aldridge, R.J. (Ed.), tasmanite oil shales from Tasmania, Australia. Geochimica et Cosmochimica Acta
Palaeobiology of Conodonts. Ellis Horwood Ltd, Chichester, pp. 129–145. 58, 3803–3822.
Jeppsson, L., 1990. An oceanic model for lithological and faunal changes tested on the Rusk, D.C., 2001. Libya: petroleum potential of the underexplored basin centers — a twenty-
Silurian record. Journal of the Geological Society of (London) 147, 633–674. first century challenge. In: Downey, M.W., Threet, J.C., Morgan, W.A. (Eds.), Petroleum
Jeppsson, L., 1993. Silurian events: the theory and the conodonts. Proceedings of the provinces of the twenty-first century. Memoir, vol. 74. American Association of
Estonian Academy of Sciences 42, 23–27. Petroleum Geologists, pp. 429–452.
Jeppsson, L., 1997. The anatomy of the Mid-Early Silurian Ireviken Event and a scenario Saltzman, M.R., 2001. Silurian δ13C stratigraphy: a view from North America. Geology
for P-S events. In: Brett, C.E., Baird, G.C. (Eds.), Palaeontological Events, 29, 671–674.
Stratigraphic, ecological and evolutionary implications. Columbia University Saltzman, M.R., 2002. carbon isotope (δ13C) stratigraphy across the Silurian-Devonian
Press, New York, pp. 451–492. transition in North America: evidence for a perturbation of the global carbon cycle.
Jeppsson, L., Talent, J.A., Mawson, R., Simpson, A.J., Andrew, A.S., Calner, M., Whitford, D.J., Palaeogeography, Palaeoclimatology, Palaeoecology 187, 83–100.
Trotter, J.A., Sandström, O., Caldon, H.-J., 2007. High-resolution Late Silurian Samtleben, C., Munnecke, A., Bickert, T., Pätzold, J., 1996. The Silurian of Gotland
correlations between Gotland, Sweden, and the Broken River region, NE Australia: (Sweden): facies interpretation based on stable isotopes in brachiopod shells.
lithologies, conodonts and isotopes. Palaeogeography, Palaeoclimatology, Palaeoecol- Geologische Rundschau 85, 278–292.
ogy 245, 115–137. Samtleben, C., Munnecke, A., Bickert, T., 2000. Development of facies and C/O-isotopes
Kaljo, D., Martma, T., 2006. Application of carbon isotope stratigraphy to dating the in transects through the Ludlow of Gotland: evidence for global and local influences
Baltic Silurian rocks. GFF 123, 123–129. on a shallow-marine environment. Facies 43, 1–38.
Kaljo, D., Kiipli, T., Martma, T., 1988. Correlation of carbon isotope events and Scotese, C.R., McKerrow, W.S., 1990. Revised world map and introduction. In: McKerrow,
environmental cyclicity in the East Baltic Silurian. In: Landing, J., Johnson, M.E. W.S., Scotese, C.R. (Eds.), Palaeozoic Palaeogeography and Biogeography. Memoir,
(Eds.), Silurian cycles – linkages of dynamic stratigraphy with atmospheric, oceanic, vol. 12. Geological Society of London, pp. 1–21.
and tectonic changes. New York State Museum Bulletin, pp. 297–312. Storch, P., Massa, D., 2007. Middle Telychian (upper Llandovery, Silurian) graptolites
Kaljo, D., Boucot, A.J., Corfield, R.M., Le Hérissé, A., Koren, T.N., Kriz, J., Männik, P., Märss, from boreholes of northwestern Libya: their biostratigraphic significance and
T., Nestor, V., Shaver, R.H., Siveter, D.J., Viira, V., 1995. Silurian bio-events. In: palaeogeographical implication. Geobios 40, 535–540.
Walliser, O.H. (Ed.), Global Events and Event Stratigraphy in the Phanerozoic. Strauss, H., Vidal, G., Moczydlowska, M., Paczesna, J., 1997. Carbon isotope geochemistry
Springer, Berlin, pp. 173–223. and palaeontology of Neoproterozoic to early Cambrian siliciclastic successions in
Kaljo, D., Kiipli, T., Martma, T., 1997. Carbon isotope event markers through the Wenlock- the East European Platform, Poland. Geological Magazine 134, 1–16.
Pridoli sequence at Ohesaare (Estonia) and Priekule (Latvia). Palaeogeography, Stricanne, L., Munnecke, A., Pross, J., 2006. Assessing mechanisms of environmental change:
Palaeoclimatology, Palaeoecology 132, 211–233. palynological signals across the Late Ludlow (Silurian) positive excursion (δ13C, δ18O) on
Kump, L.R., Arthur, M.A., Patzkowsky, M.E., Gibbs, M.T., Pinkus, D.S., Sheehan, P.M., 1999. Gotland, Sweden. Palaeogeography, Palaeoclimatology, Palaeoecology 230, 1–31.
A weathering hypothesis for glaciation at high atmospheric pCO2 during the Late Talent, J.A., Mawson, R., Andrew, A.S., Hamilton, P.J., Whitford, D.J., 1993. Middle
Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology 152, 173–187. Palaeozoic extinction events: faunal and isotopic data. Palaeogeography, Palaeocli-
Lécuyer, C., Paris, F., 1997. Variability in the δ13C of lower Palaeozoic palynomorphs: matology, Palaeoecology 104, 139–152.
implications for the interpretation of ancient marine sediments. Chemical Geology Tongiorgi, M., Yin, L.-M., Di Milia, A., 2003. Lower Yushanian to lower Zhejiangian
138, 161–170. palynology of the Yangtze Gorges area (Daping and Huanghuachang sections),
Legrand, P., 1985. Lower Palaeozoic rocks of Algeria. In: Holland, C.H. (Ed.), Lower Hubei province, South China. Palaeontographica. Abteilung B 266, 1–160.
Palaeozoic of north-western and west central Africa, Lower Palaeozoic rocks of the Tyson, R.V., 1995. Sedimentary organic matter. Chapman & Hall, London. 616 pp.
world, vol. 4. John Wiley and Sons Ltd., New York, pp. 5–89. Tyson, R.V., 2006. Calibration of hydrogen indices with microscopy: A review, reanalysis
Lehnert, O., Fryda, J., Buggish, W., Munnecke, A., Nützel, A., Kriz, J., Manda, S., 2007. δ13C and new results using the fluorescence scale. Organic Geochemistry 37, 45–63.
records across the late Silurian Lau event: new data from middle palaeo-latitudes of Vecoli, M., 2000. Palaeoenvironmental interpretation of microphytoplankton diversity
northern peri-Gondwana (Prague Basin, Czech Republic). Palaeogeography, trends in the Cambrian–Ordovician of the northern Sahara Platform. Palaeogeo-
Palaeoclimatology, Palaeoecology 245, 227–244. graphy, Palaeoclimatology, Palaeoecology 160, 329–346.
Loydell, D.K., 2007. Early Silurian positive δ13C excursions and their relationship to Vecoli, M., 2008. Fossil microphytoplankton dynamics across the Ordovician–Silurian
glaciations, sea-level changes, and extinction events. Geological Journal 42, 531–546. boundary. Review of Palaeobotany and Palynology 148, 91–107.
Lüning, S., Craig, J., Loydell, D.K., Storch, P., Fitches, B., 2000. Lower Silurian “hot shales” Vecoli, M., Le Hérissé, A., 2004. Biostratigraphy taxonomic diversity, and patterns of
in North Africa and Arabia: regional distribution and depositional model. Earth- morphological evolution of Ordovician acritarchs (organic-walled microphyto-
Science Reviews 49, 121–200. plankton) from the northern Gondwana margin in relation to palaeoclimatic and
Lüning, S., Loydell, D.K., Storch, P., Shahin, Y., Craig, J., 2006. Origin, sequence palaeogeographic changes. Earth-Science Reviews 67, 267–311.
stratigraphy and depositional environment of an Upper Ordovician (Hirnantian) Verniers, J., Nestor, V., Paris, F., Dufka, P., Sutherland, S., Van Grootel, G., 1995. A global
deglacial black shale, Jordan — Discussion. Palaeogeography, Palaeoclimatology, Chitinozoa biozonation for the Silurian. Geological Magazine 132, 651–666.
Palaeoecology 230, 352–355.
Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

Contents lists available at ScienceDirect

Palaeogeography, Palaeoclimatology, Palaeoecology


j o u r n a l h o m e p a g e : w w w. e l s ev i e r. c o m / l o c a t e / p a l a e o

Organic-carbon deposition and coastal upwelling at mid-latitude during the Upper


Ordovician (Late Katian): A case study from the Welsh Basin, UK
T.J. Challands a,⁎, H.A. Armstrong a, D.P. Maloney a, J.R. Davies b, D. Wilson b, A.W. Owen c
a
Department of Earth Sciences, Durham University, Science Laboratories, South Road, Durham, DH1 3LE, UK
b
British Geological Survey, Kingsley Dunham Centre, Keyworth, Nottingham, NG12 5GG UK
c
Department of Geographical Earth Sciences, University of Glasgow, Gregory Building, Lilybank Gardens, Glasgow, G12 8QQ, UK

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

Article history: A lack of stratigraphical, sedimentological and geochemical data for sediment accumulation rates and
Received 2 October 2007 indicators of productivity and anoxia means that causative models for ancient black shales are largely
Received in revised form 19 July 2008 inferred from modern settings. Coastal upwelling has been suggested as a general hypothesis for Ordovician
Accepted 9 October 2008
black shale deposition within the Iapetus Ocean, but has not been directly tested. Despite anchizone
metamorphism we utilize a suite of geological and geochemical environmental proxy data (TOC wt.%, δ13Corg,
Keywords:
Ashgill
Ba/Al, P) to elucidate the mechanism for the origin of a single grey-black shale cycle within the upper Katian
Upper Ordovician succession of the Welsh Basin. Here we interpret organic carbon (OC)-rich deposition to be suggestive of a
Katian period of high photic productivity (higher TOC wt.%), with 12C and Ba enrichment, comparable to high
Organic carbon productivity events in modern coastal upwelling systems. Productivity proxy values are consistent with those
δ13Corg from margins where sedimentation rates are high (e.g. Gulf of California). Inter-bedded grey shales have low
Trace metal TOC wt.% more positive δ13Corg and marginally lower Ba and are interpreted as low productivity events.
Redox Thalassinoides, Planolites and Chondrites ichnofacies indicate changing seafloor oxygen levels. These show
Upwelling
predominantly dysoxic conditions even during the deposition of OC-poor grey shales. During an OC-rich
TOC
laminated hemipelagite event, oxygen levels declined at the sea floor before the deposition of the OC-rich
Climate
layers. Full anoxia was established early on in the deposition of OC-rich layers and the return to more oxic
conditions was rapid though fluctuating and coincident with the return of grey shale deposition. This pattern
suggests OC accumulation at the seafloor resulted from a complex interaction of productivity and
preservation. None of the widely used trace element redox proxies are reliable in anchizone metamorphic
rocks. Climate sensitive detrital proxies (K/Al and Ti/Al) indicate arid–temperate conditions in the basin
hinterland during the deposition of the OC-rich layers. The Welsh Basin was situated on the southern margin
of the Iapetus Ocean in (30°S) beneath the prevailing SE trade winds. We interpret the occurrence of OC-rich
laminated hemipelagite events to represent the intensification of upwelling, prior to the Hirnantian
glaciation, possibly having resulted from a strengthening of the trade winds, associated with stepped changes
in ice volume, and a more arid local climate.
© 2008 Published by Elsevier B.V.

1. Introduction climatic conditions it is necessary to determine whether the processes


that formed Mesozoic to modern ocean black shales are also
Black shales provide important records of climate and atmo- applicable to Palaeozoic black shales. Processes for black shale
sphere–ocean interactions. The Early Palaeozoic oceans differed from formation can typically be considered to be the product of increased
the Mesozoic and younger oceans because they experienced a mild productivity or enhanced preservation (see Tyson, 1995, for a review).
greenhouse climate (Yapp and Poths, 1992; Berner, 1994; Berner and Increasing productivity has been explained by: 1) increasing con-
Kothavala, 2001) and were prone to permanent anoxia (Leggett, 1980). tinental runoff (the Runoff model e.g. Beckmann et al., 2005; Bjerrum
However, black shales deposited in the Early Palaeozoic potentially et al., 2006), dominant at tropical latitudes, whereby increased
provide a detailed record of climate processes provided the mechan- continental runoff introduces nutrients, increases productivity and
isms for their formation can be constrained. Under such different organic carbon (OC) export. Weathering regime proxies (K/Al, Ti/Al)
have recently demonstrated that fresh-water and detrital input into
basins is important for the formation of oxygen minima and the burial
⁎ Corresponding author. Current address: Total E & P UK Limited, Crawpeel Road,
of carbon e.g. the Cretaceous Ivorian Basin (Beckmann et al., 2005;
Altens Industrial Estate, Aberdeen, AB12 3FG, UK. Bjerrum et al., 2006; Sobarzo et al., 2007). Along the upwelling margin
E-mail address: tom.challands@total.com (T.J. Challands). of Chile, high freshwater input suppresses upwelling through the

0031-0182/$ – see front matter © 2008 Published by Elsevier B.V.


doi:10.1016/j.palaeo.2008.10.004
396 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

formation of a salinity stratified layer during wetter periods resulting et al., 2007) but does not explain variation in productivity within black
in lower oxygen values at depth (Sobarzo et al., 2007). 2) increased shale successions, and in fact assumes it to be effectively constant
oceanic/coastal upwelling (the Upwelling model e.g. Parrish, 1982; (Berry and Wilde, 1978, p. 271; Leggett, 1980, p. 150). For example, the
Hay and Brock, 1992; Pope and Steffen, 2003) dominant at mid- to upper Katian (Rawtheyan Hirnantian British Stages) succession of
high-latitudes in trade wind zones. Mechanisms for increasing Dob's Linn, Scotland, located on the northern margin of the Iapetus
preservation include: 1) the Transgressive model, whereby transgres- Ocean contains thin (b30 cm thick) black shale units alternating with
sion introduces additional nutrients and expansion of the deep ocean grey shale. These have been inferred to have been deposited at a time
oxygen minimum zone (OMZ) during sealevel rise promotes OC of increased productivity (Armstrong and Coe, 1997) and global
preservation in high productivity shelf environments (Leggett, 1980), eustatic, glacially-induced regression when the Iapetus Ocean was
2) restricted circulation, either global thermohaline circulation or much narrower and likely more restricted than has been previously
regional circulation such as in a restricted basin, and, 3) increased considered (Armstrong and Owen, 2002). Two oceanic events have
burial efficiency of OC during increased sedimentation rate (Canfield, been recognized in the Dob's Linn succession (Fig. 1). The first, in the
1989). mid Katian (Streffordian–Pusgillian boundary), is the sudden change
The Transgressive model with unrestricted oceans may explain from anoxic to oxic conditions and the replacement of black with grey
thick black shales in Lower Palaeozoic basins (Leggett, 1980; Page shales. This has been interpreted as the onset of glacially-induced

Fig. 1. Stratigraphic chart (a) showing possible correlation between the sections at Dob's Linn (international boundary stratotype for the base of the Silurian System), lithostratigraphy
and graptolite biostratigraphy for Dob's Linn after Melchin et al. (2003); Underwood et al. (1997); Armstrong and Coe (1997). (b) Palaeogeographic map of Wales, England and
southern Scotland during the Late Ordovician showing inferred palaeobathymetry (based on Woodcock and Strachan, 2000). (c) Palaeogeographic reconstruction of Avalonia in the
Upper Ordovician showing position of the Welsh Basin (based on Cocks et al., 1997). (d) Palaeogeographic map of the Iapetus Ocean during the Late Ordovician (based on Herrmann
et al., 2005). This shows the distribution of the major continental land masses (L = Laurentia, S = Siberia, B = Baltica, A = Avalonia, DL = Dob's Linn, WB = Welsh Basin) and major surface
currents (solid arrows).
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 397

thermohaline circulation (Armstrong and Coe, 1997). The second is the deposition from proxy data. Coastal upwelling is the preferred cause
re-appearance of discrete black shale units within predominantly grey of organic carbon (OC) enrichment of the sediments, and could have
shales during the upper Katian (anceps graptolite Biozone), and been driven by changes in the strength of the prevailing trade winds.
was considered indicative of climatically-forced, Monterey-type Given the limitations of the extent and fidelity of the data, other
processes reflecting upwelling (Armstrong and Coe, 1997). The possible mechanisms are: 1) increased anoxia and, 2) marine
presence of similar lithostratigraphic events in the Welsh Basin are transgression. The upwelling hypothesis, however, is consistent with
potentially contemporaneous with those at Dob's Linn and may known patterns of Late Ordovician climate change associated with the
demonstrate that they could have been Iapetus Oceanwide reflecting developing Hirnantian glaciation.
fundamental changes in the climate–ocean system before the onset of
glaciation. 2. Regional geology of the Ordovician of the Welsh Basin
Oceanic upwelling has been inferred from lower Katian (mid-
Caradoc) radiolarian cherts along the Iapetus Ocean margin from Two major changes in oceanography can be identified in the Welsh
Laurentia (Pope and Steffen, 2003) and in Wales from the presence of Basin during: 1) the mid-Katian (late Caradoc to early Ashgill) in the
phosphorites and radiolarites (e.g. Cave, 1965). The latter were linearis complanatus graptolite biozones (spinifera chitinozoan Bio-
compared to phosphorites forming today at shelf-slope depths off zone) and, 2) in the late Katian anceps graptolite Biozone (umbilicata–
the California coast and attributed to climatic factors reflecting an gamachiana chitinozoan biozones). At the first event, anoxic condi-
absence of currents and strong offshore winds or basin restriction tions were terminated and the succession changes to predominantly
from the presence of a bathymetric barrier (Cave, 1965). burrow-mottled lithofacies. These burrow-mottled grey shale units
In this case study we use a suite of proxy data to test an upwelling have been mapped regionally as the Nantmel Mudstones Formation
mechanism for an upper Katian black shale unit within the Welsh (Davies et al., 2003; Fig. 2). During the late Katian a return to black
Basin. Despite anchizone facies metamorphism, it has been possible to shale deposition is recognized (the “Red Vein”) comprising distinctive
reconstruct plausible palaeo-environmental conditions during alternating black laminated and grey burrow mottled siltstone and

Fig. 2. Regional Geology of the late-Katian of the Cardigan–Llangranog region, southwest Wales, UK. Abbreviations on regional location map BW = Builth Wells, CI = Cadair Idris,
LL = Llangranog, NE = Newcastle Emlyn, Rhay = Rhayader. Chronostratigraphy from Fortey et al. (2000) and Webby et al. (2004), chitinozoan biozonation from Vandenbroucke and
Vanmeirhaeghe (2007), graptolite biozonation and Time slices from Webby et al. (2004), lithostratigraphy from Davies et al. (2003).
398 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

mudstones. The cause of this later change in lithofacies is unknown places and sections that allow safe access to the LH units and
and is the focus of this paper. interbedded OC-poor lithofacies have been selected: LH0 is accessible
The Red Vein (Fig. 2) crops out from the Cadair Idris region in the at Aberporth (SN 625 515), the base of LH1 is exposed at Tresaith (SN
north, south to Rhayader and Builth Wells regions and further west 280 516) and the upper boundary of LH1 crops out and is easily
around Newcastle Emlyn and north of Cardigan (British Geological reached at Traeth Penbryn (SN 289 522) (see Fig. 2). The same sections
Survey Llangranog 1:50 000 scale Sheet number 194; 2003). This for trace fossil analysis and geochemical analysis could not be used for
distinctive unit was first recognized in the Cadair Idris area by Pugh reasons of access. All trace fossil data were collected from LH0 at
(1923) and named on account of oxidation of the abundant iron Aberporth (SN 625 515, Locality 1), two sections spanning the lower
sulphide it contains. Three separate OC-rich laminated units have and upper parts of LH0 are conflated to a single composite section
previously been recognized in the Rhayader, Builth Wells and (Fig. 5). This 6.4 m-thick section was sub-sampled over six 40 cm
Llangranog regions (Davies et al., 1997; Schofield et al., 2004; Davies intervals, each separated by approximately 1 m.
et al., 2006) and interpreted as laminated hemipelagites (LH, Davies Geochemical samples were collected from two 5 m thick sections
et al., 1997). They consist of thin alternating laminae of dark grey, OC- spanning both the upper and lower boundaries of LH1. The lower
rich and medium grey OC-poor mudstone. boundary was collected at Tresaith (SN 280 516) and the upper
The LH units in the Llangranog region range in thickness from boundary at Traeth Penbryn (SN 289 522, Localities 2 and 3 respectively,
between 1036 m. Laminated OC-rich layers after 1.4 m above the base of see Fig. 2). Samples for trace element analysis were taken through LH1
LH0 contain abundant pyrite framboids and small nodules up to 1 mm in from organic-rich horizons at 20 cm intervals. There is no evidence at
length alongside finely-disseminated pyrite. The OC-rich laminated the sampling sites of secondary mineralization and all samples were
horizons are composed of a fine clay matrix and angular quartz grains up prepared from specimens with no surface alteration or fracturing.
to 0.25 mm and fine, platy muscovite and chlorite grains up to 0.175 mm Powders were then drilled from samples using a tungsten carbide drill
long. The medium grey OC-poor mudstone component of the LH units bit to homogenize and reduce matrix effects. Matrix effects were
occasionally contains thin fining-up sequences from silt to mudstone considered negligible due to the fine-grained nature of the samples.
and occasional very thin (b2 mm) fine-grained sand and are interpreted Sub-samples were taken for trace element and stable isotopic analysis.
as hemiturbidites (Stow and Wetzl, 1987) deposited in a slope-apron Sample digestion was carried out using HF-HN03 standard acid
system (Schofield et al., 2004). Within the LH, the absence of significant digestion techniques for ICPMS following the procedure of Ottley
grain-size variations between the OC-poor grey mudstones and OC-rich et al. (2003). A Perkin Elmer-Sciex Elan 6000 ICP-MS used in
units suggest the increase in OC content of the sediment was conjunction with a Cetac Direct Injection Nebuliser and a Cetac Aridius
independent of factors affecting sediment supply though precise Desolvating Nebuliser was used for analysis of trace elements. The
measures of sedimentation rate in ancient basin settings are difficult machine was calibrated using international rock standards W2, BHV01,
to measure (Wignall, 1994). AGV1, Be-N, NBS688 and BIR-1 and shale standards ScO-1, MAG-1 and
On regional maps of these areas the laminated hemipelagite units SGR-1 were also used for initial comparison of sample shale values. All
are termed lh'–lh″′ (British Geological Survey Rhayader 1:50 000 scale samples were analysed over three separate runs and reproducibility
Sheet numbers 178 and 179; 1997, British Geological Survey Builth throughout runs was monitored using SGR-1 (see Table 1). Relative
Wells 1:50 000 scale Sheet number 196; 2004, British Geological standard deviation (σ / mean× 100) was less than 11% in all runs.
Survey Llangranog 1:50 000 scale Sheet number 194; 2006). Herein, lh Internal precision for individual analyses was greater than 95%.
are capitalized and referred to numerically (e.g. LH1) to avoid confusion Normalisation to Al is a standard procedure for comparing element
between the letter ‘l’ and the number ‘1’ and to prevent inconsistency proportions, particularly where organic matter content is variable (e.g.,
in labelling. In the Llangranog region, an older LH horizon has been Tribovillard et al., 2005). Certain elements e.g. Ni are traditionally
recognized in this study and, following the British Geological Survey plotted against other detrital indicators e.g. Co. Spot samples for δ13Corg
numbering convention for these units, has been named LH0. and TOC wt.% were also collected from eleven other localities
The LH units are interbedded with grey burrow-mottled fine- representative of OC-poor facies below, between and above LH units
grained sandstones, siltstones and mudstones up to 25 m thick that and from LH0, LH2 and LH3 (Table 2, Fig. 4). All δ13Corg and TOC %
have been interpreted as fine grained distal turbidites resulting from measurements were made on a Thermo Finnigan MAT 253 Stable
periodic influx of coarser material (Davies et al., 2003; Schofield et al., Isotope Mass spectrometer at Durham University, Department of Earth
2004). They are hereafter referred to the OC-poor lithofacies. Sciences. TOC wt.% analysis of samples from 779.5–794.5 cm from the
Phosphatic nodules up to 1 m in length and phosphate horizons up to base of LH1 were conducted at Newcastle University Department of Civil
10 cm thick are present in both the organic-rich and organic-poor facies. Engineering and Geosciences using Rock Eval (6) Pyrolysis. Urea2, CH-7
Diffuse fine-grained phosphatic layers up to 10 mm thick are more and CH-6 standards were used for machine calibration and standard
frequent in the LH units and less common in the OC-poor lithofacies. In
the OC-poor lithofacies they occur in the light-grey silty mud layers.
Table 1
The thermal history of the Ordovician of the Welsh Basin has been
Reproducibility and error data for all shale ICP-MS runs using shale standard SGR-1
elucidated from illite crystallinity (Fettes et al., 1985; Robinson and
Bevins,1986; Bevins and Robinson,1988; Roberts et al.,1996). These data Run-1, n = 2 Run-2, n = 7 Run-3, n = 2
indicate a regional record of diagenetic to epizone facies metamorphism. Mean StDev RSD% Mean StDev RSD% Mean StDev RSD%
The Nantmel Mudstones Formation adjacent to the study area (British Al 6.1 0.15 2.47 6.92 0.5 7.2 18.66 1.37 7.36
Geological Survey 1:50 000 Sheet 193, Cardigan and Dinas Island) is of K 1.57 0.03 1.71 1.63 0.07 4.48 2.24 0.0 0.22
Ti 0.24 0.01 3.85 0.24 0.01 2.60 0.26 0.02 7.18
low anchizone grade. Darriwillian lower Katian aged shales immediately
V 120.56 3.88 3.22 121.82 2.70 2.22 124.12 7.81 6.29
south of Cardigan, have experienced late diagenetic to epizone Cr 29.62 0.34 1.16 33.31 0.95 2.85 33.60 1.87 5.56
temperatures (200–300 °C Lev et al., 2008) indicating burial to c. 4– Mn 0.03 0.00 0.00 0.03 0.00 2.43 0.03 0.00 4.29
6 km using a regional geothermal gradient of c. 50 °C/km (Bevins and Co 11.35 0.29 2.52 11.87 0.19 1.63 12.39 0.39 3.12
Ni 25.69 0.52 2.01 31.10 0.70 2.25 31.62 1.23 3.90
Merriman, 1988; Bottrell et al., 1990; Roberts et al., 1991).
Cu 60.60 0.44 0.72 60.73 0.53 0.87 63.84 1.91 3.00
Ba 299.63 1.89 0.63 295.20 8.16 2.77 263.07 8.72 3.32
3. Materials and methods Th 4.65 0.14 3.04 4.61 0.12 2.50 4.38 0.24 5.56
U 5.17 0.03 0.67 5.09 0.11 2.14 4.68 0.28 6.05
The best sections through the Red Vein are exposed in cliff sections Each run included both OC-rich and OC-poor samples. Al, K, Ti and Mn in wt.% oxide, all
along the coast of Cardigan Bay (Fig. 2). Accessibility is difficult in other elements in ppm.
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 399

Table 2 elements (Cu, Cr, Ni, V, Cd, Mo and U Tribovillard et al., 2005). These
δl3Corg and TOC for Cardigan coast line, Ceredigion, Wales elements form sulphides and complex with organic matter under
Sample Locality δ13Corg ‰ TOC % reducing conditions. OC-rich sediments are therefore typically
no. enriched in these metals (De La Rocha, 2004; Tribovillard et al.,
1 Gwbert Hotel (SN 161 500) – 29.46 0.29 2005, and references therein).
2 Carreg Lydan (SN 162 513) – 28.1 0.28 However some of these elements show significant mobility during
3 Mwnt (SN 192 519) – 30.95 0.27
late diagenesis to low grade metamorphic conditions. For example, at
4 Pen-y-Craig (SN 218 523) – 28.97 0.08
5 Aberporth (SN 258 515) – 30.70 0.15 these grades V becomes mobile and complexes with illite (Peacor et al.,
6 LH0 lower boundary, subsection 3 (SN 625 515) – 30.91 0.17 2000) rendering the frequently used redox proxies V/(V + Ni) and V/Cr
7 LH0 2 m, subsection 3 (SN 625 515) – 31.83 0.28 as unreliable. Uranium is also mobilized and enriched, whilst Th is
8 LH0 upper boundary, subsection 6 (SN 625 515) – 30.98 0.16 depleted during the formation of illite at low metamorphic grades
9 LH1 −80 cm from base of LH1, Tresaith (SN 280 516) – 30.5 0.14
10 LH1 −40 cm from base of LH1, Tresaith (SN 280 516) – 31.16 0.53
(Hannigan and Basu, 1998). Nickel, however, has been regarded as a
11 LH1 280 cm from base of LH1, Tresaith (SN 280 516) – 31.47 0.38 reliable redox proxy in organic-rich shales (Jones and Manning, 1994)
12 LH1 320 cm from base of LH1, Tresaith (SN 280 516) – 31.71 0.36 and has not been demonstrated to be mobile during low metamorph-
13 LH1 360 cm from base of LH1, Tresaith (SN 280 516) – 31.51 0.51 ism. We therefore use Ni/Co (Dypvik, 1984) and Ni/Al (Tribovillard
14 LH1 400 cm from base of LH1, Tresaith (SN 280 516) – 31.78 0.45
et al., 2005) ratios as an indicator of redox conditions in this study.
15 LH1 440 cm from base of LH1, Tresaith (SN 280 516) – 31.42 0.34
16 LH1 480 cm from base of LH1, Tresaith (SN 280 516) – 31.77 0.45
17 LH1 520 cm from base of LH1, Tresaith (SN 280 516) – 31.48 0.40 3.2. Productivity proxies
18 LH1 560 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.48 0.39
19 LH1 600 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.73 0.42 3.2.1. δ13Corg
20 LH1 640 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.53 0.45
The δ13Corg values of sedimentary organic matter are controlled by a
21 LH1 680 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.81 0.44
22 LH1 720 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.69 0.47 number of factors including organic matter composition and cell growth
23 LH1 760 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.12 0.29 rates (for a review see Freeman, 2001; Maslin and Swann, 2005, and
24 LH1 800 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.77 0.51 references therein). In the open ocean, sedimentary organic matter is
25 LH1 822.9 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.61
derived from marine and terrestrial sources. In pre-Devonian time there
26 LH1 823.5 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.61
27 LH1 824 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.524 was no extensive terrestrial vegetation (Peters-Kottig et al., 2006) and
28 LH1 824.8 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.255 organic matter in Ordovician rocks can therefore be considered entirely
29 LH1 825.9 cm from base of LH1, Traeth Penbryn (SN 289 522) – 31.52 0.258 marine in origin. Cell growth rates cannot be constrained in ancient
30 LH1 826.4 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.522 samples. Bulk sediment δ13Corg values are therefore only considered
31 LH1 826.9 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.281
indicative of shifts in mean oceanic isotopic composition.
33 LH1 827.5 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.604
33 LH1 828 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.469 δ13Corg values have been shown to yield an unaltered environ-
34 LH1 830 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.322 mental signal in greenschist facies rocks at Dob's Linn (Underwood
35 LH1 830.5 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.226 et al., 1997) and are therefore considered primary oceanic values in
36 LH1 831.6 cm from base of LH1, Traeth Penbryn (SN 289 522) – 30.3 0.324
this study. Nineteen δ13Corg samples were taken through the LH1. One
37 LH1 832.5 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.267
38 LH1 833.3 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.255
whole rock sample from each of the remaining three laminated
39 LH1 833.8 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.313 hemipelagite units (LH0, LH2 and LH3) was taken for δ13Corg as well as
40 LH1 835.3 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.3 nine samples taken from OC-poor lithofacies below, between and
41 LH1 836.9 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.275 above the LH units (Table 2).
42 LH1 837.9 cm from base of LH1, Traeth Penbryn (SN 289 522) – 0.255
43 LH1 842.4 cm from base of LH1, Traeth Penbryn (SN 289 522) – 29.8 0.2
44 LH2 (SN 293 523) – 31.73 0.41 3.2.2. Total organic carbon (TOC wt.%)
45 LH3 lower boundary, org.-poor facies (SN 296 528) – 30.8 0.44 TOC wt.% is known to correlate with more direct measures of
46 LH3 lower boundary org.-rich facies (SN 296 528) – 31.87 0.40 photosynthetic primary productivity such as total chlorophyll-a or
47 Highest Nantmel Mudstones Formation (SN 296 530) – 30.50 0.24
total steryl chlorine esters (Nara et al., 2005) and organic mass
Samples are listed from stratigraphically lowest (oldest) to highest (youngest), see Fig. 4 accumulation rate (Tyson, 1995; Vilinski and Domack, 1998; Kuypers
for stratigraphical plot of this data. et al., 2002, 2004; Twichell and Diester-Haass, 2002; Meyers and
Arnaboldi, 2005) and may therefore be suitable as a proxy for
deviation was less than 0.15‰ for all standards. Carbon isotope ratios productivity when supported by other proxy data (e.g. Ba). The burial
were expressed in the standard delta (δ) notation in per mil (‰) against efficiency of carbon is controlled by oxygen level and burial rate (for
the internationally accepted standard notation, Vienna Peedee Belem- example, Wignall, 1994; Tyson, 2001) and is therefore only likely to be
nite (VPDB). Analytical reproducibility of replicate samples using this representative of productivity when these two factors can unambigu-
method was better than 0.4‰. ously be shown to be constant. Thirty-five TOC samples were
measured through the OC-poor LH1 OC-poor section at Tresaith-
3.1. Redox proxies Penbryn and, like δ13Corg, one whole rock sample from each of the
remaining three laminated hemipelagite units (LH0, LH2 and LH3) was
3.1.1. Trace fossils measured for TOC wt.% as well as nine samples taken from OC-poor
The application of analysis of trace fossil assemblages and the degree facies below, between and above the LH units (Table 2).
of bioturbation has proven a strong tool for determining fluctuations in
sediment and basin oxygenation e.g. Ekdale and Mason (1988), Savrda 3.2.3. Trace elements
and Bottjer (1994), Savrda (1995), Löwemark et al. (2004). Ichnogenera The strong correlation between Ba and organic matter in marine
presence–absence, burrow diameter and bioturbation index data were sediments has led Ba values to be used to infer past export production
collected in line with standard procedures (Savrda and Bottjer, 1994; (De La Rocha, 2004). Dysoxic conditions are characterized by relatively
Savrda, 1995; Löwemark et al., 2004) at centimetre-resolution. low Ba concentrations (b500 ppm), increased Cd and U (Prakash Babu
et al., 2002). Though frequently used as a palaeoproductivity indicator,
3.1.2. Trace elements Ba has been found to be associated with sapropels formed in euxinic
The biological pump heavily influences the cycling, concentration basins (Brumsack, 2006). In these environments, Ba is highly mobile
and residence times of the redox-sensitive and sulphide forming trace and precipitates as barium-sulphides. Ba is thus an unreliable
400
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410
Fig. 3. Geochemical proxy data plotted against stratigraphical height above base of LHl, Traeth Penbryn. Error bars represent 1−σ.)
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 401

Fig. 4. Composite stratigraphical chart showing variations in δ13Corg and TOC % profile throughout the upper Katian (Cautleyan–Rawtheyan, Upper Ordovician) of the Cardigan–
Penbryn coastal section, Welsh Basin (SN 161 500 to SN 296–530). See Table 3 for details of precise sample locations.

indicator of productivity if euxinia can be demonstrated (Shimmield, metamorphic rocks we have normalized to Al which is also considered
1992; Prakash Babu et al., 2002). Ba is typically measured as a to be a detrital indicator (Tribovillard et al., 2005). We also plot raw
ratio against Th which is taken to represent terrigenous detrital Ba abundance (ppm) to test if an increase in Ba/Al is the product of
input (Shimmield, 1992). Due to the mobility of Th in low grade decreasing Al.

Table 4a
Table 3 f and t-test statistics for K/Al and Ti/Al through LH1, Traeth Penbryn
f and t-test statistics for Ba/Al and Ba (ppm) through LH1, Traeth Penbryn
OC-poor LH1 t-stat t-crit OC-poor LH1 f-stat f crit
OC-poor LH1 t-stat t-crit OC-poor LH1 f-stat f crit Mean Mean 2-tail Var. Var. 1-tail
Mean Mean 2-tail Var. Var. 1-tail K/Al 0.16 0.17 0.45 2.03 0.3 × 10− 3 1.92 × 10− 5 0.06 0.6
Ba/Al 2.7 × 10− 3 2.8 × 10− 3 0.6 2.02 1.69 × 10− 7 2.08 × 10− 8 8.12 1.74 Ti/Al 0.046 0.043 −3.77 2.03 2.1 × 10− 5 1.7 × 10− 6 0.08 0.57
Ba (ppm) 569.8 671.9 4.48 2.03 15878.6 866.6 18.32 1.75
Both the means and variances for K/Al are significantly different at α = 0.05 between OC-
Both the means and variances for Ba/Al and Ba (ppm) are significantly different at poor and OC-rich units. For Ti/Al, only the variance is significantly at α = 0.05 between
α = 0.05 between OC-poor and OC-rich units. OC-poor and OC-rich units is different.
402 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

Table 4b 3.3. Climate sensitive detrital input proxies, K/Al and Ti/Al
Trace element data, LH1, Traeth Penbryn

Depth (cm) Al K Ti Co Ni Ba K and Al are mainly confined to the fine-grained aluminosilicate


1040 20.30 2.55 0.78 30.38 85.26 398.42 fraction, clay minerals and feldspars. Kaolinite which is rich in Al
1020 21.48 3.30 0.97 20.15 57.48 523.08 forms by intense chemical weathering under humid conditions
1000 22.57 3.54 1.00 21.41 61.18 558.80 whereas illites, which are rich in K, form under less humid
980 22.95 3.78 1.02 21.08 53.14 599.76
weathering conditions (Beckmann et al., 2005; Tribovillard et al.,
960 20.38 2.37 0.72 33.00 87.89 376.33
940 22.70 3.65 1.05 11.00 42.86 635.11 2005). Ti is concentrated in heavy minerals and is often transported
920 22.90 3.64 1.04 17.49 47.34 575.91 as part of the wind blown detritus, considered indicative of
900 20.94 3.18 0.86 18.60 51.35 566.61 weathering in arid environments. The K/Al ratios are therefore
880 22.98 3.74 1.05 28.99 55.59 500.10 used as a proxy of temperate to humid weathering and, similarly, Ti/Al
860 23.62 4.05 1.10 20.63 51.74 582.61
840 22.20 3.93 1.01 10.03 42.02 636.98
834.5 20.46 3.77 0.90 8.31 29.55 603.73
833.5 22.31 3.87 1.08 11.18 41.09 645.21
832 22.91 4.14 1.11 7.00 46.88 677.80
830.5 22.65 3.79 1.03 8.59 33.52 658.14
829.75 21.86 3.62 0.99 8.18 35.54 641.16
828.25 17.02 3.60 1.07 5.48 28.45 674.28
826.5 19.19 3.40 1.01 5.47 26.77 719.64
824 21.85 3.71 1.08 8.01 27.05 713.92
823 22.11 4.06 1.10 6.80 29.65 615.95
822 10.31 1.90 0.53 5.85 10.32 293.64
820.75 7.84 1.41 0.40 9.41 14.24 230.62
819.5 9.14 1.60 0.43 40.09 54.74 265.87
800 22.30 3.77 1.00 49.61 92.95 629.64
780 22.87 3.74 1.00 44.51 81.83 624.06
760 23.67 3.84 1.06 25.08 68.19 597.80
740 22.83 3.75 1.00 25.56 65.33 626.40
720 23.10 3.76 1.06 27.61 74.06 624.73
700 24.14 3.98 1.03 23.87 64.82 666.42
680 24.13 3.98 1.02 24.02 65.15 667.03
660 24.57 4.08 1.05 27.66 73.52 695.30
640 24.03 3.88 1.00 26.37 69.57 650.20
620 25.41 4.21 1.09 23.45 56.92 690.73
600 24.51 4.01 1.04 25.36 70.33 668.63
580 24.70 4.12 1.06 26.79 65.52 695.13
560 25.09 4.10 1.05 21.28 56.08 674.47
540 23.73 3.92 1.06 26.13 65.38 658.18
520 23.97 3.96 1.06 22.87 62.66 674.32
500 24.25 3.90 1.05 23.69 67.88 648.63
480 24.74 4.05 1.08 27.24 66.99 674.07
460 24.89 4.09 1.05 23.53 63.67 687.62
440 25.28 4.10 1.07 22.01 60.39 677.77
420 25.30 3.94 1.05 20.51 59.14 665.27
400 26.06 4.24 1.07 21.12 65.18 709.96
380 25.31 3.95 1.03 16.86 54.90 676.80
360 24.55 3.96 1.08 17.04 56.47 675.72
340 23.31 3.70 0.95 19.84 54.56 625.34
320 24.36 3.86 1.04 21.68 60.80 660.32
300 24.00 4.00 1.03 20.92 49.83 675.16
280 25.50 4.21 1.11 19.22 53.81 694.27
260 24.38 4.07 1.03 18.90 50.77 694.41
240 23.65 3.83 1.00 17.61 52.36 648.93
220 27.61 4.53 1.15 21.29 58.55 757.73
200 23.94 4.00 0.98 24.96 56.25 700.33
180 22.83 3.97 1.01 32.36 64.47 704.44
160 23.51 4.04 1.01 32.18 63.07 717.06
140 22.41 3.76 1.00 29.52 64.09 681.81
120 22.15 3.83 0.97 26.24 56.33 679.47
100 22.03 3.77 0.99 34.86 62.89 670.49
80 22.23 3.82 0.99 24.67 56.95 682.50
60 22.52 3.85 1.01 39.03 60.55 677.27
40 22.32 3.83 0.99 41.11 62.00 675.94
20 22.38 3.80 0.97 53.31 55.01 671.13
−40 22.63 3.28 0.99 21.23 60.44 572.60
−60 23.57 3.57 1.05 20.54 56.13 629.54
−80 22.67 3.64 0.98 17.79 48.51 640.86
−100 22.58 3.56 1.00 18.84 48.24 620.59
−120 23.38 3.81 1.10 18.65 45.47 663.89
−140 23.02 3.25 1.05 24.11 65.62 584.37
−160 22.64 3.64 1.06 18.56 45.67 649.10
−180 23.21 3.68 1.09 21.44 56.50 659.44
−200 20.26 2.98 0.90 22.06 53.96 520.37

Al, K, Ti are recorded as wt % oxide, Co, Ni and Ba are recorded as ppm.


All depths from the base of LH1.

Fig. 5. Cross-plots for a) TOC % vs δ13Corg, b) Ba vs. Al and c) Ba/Al vs. TOC %.
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 403

Fig. 6. Ichnofaunal responses to organic-rich sedimentation and development of anoxia at OC-rich unit LH0, Aberporth. See text in section 3.3 for details of sampling location.
404 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

For LHl at Tresaith, the mean δ13Corg value is 31.6‰ (n = 14) and in
the OC-poor units, the δ13Corg mean is 30.7‰ (n = 5).
The more limited data from the other LH units suggest the LH
lithofacies is associated with a negative δ13Corg excursion of
approximately 1‰ (Fig. 4). δ13Corg values for each LH unit are between
30.5‰ (LHl) and 31.9‰ (LH3, mean for all LH units = 31.55‰) compared
to a mean value of 30.25‰ for all OC-poor samples.
δ13Corg values from the OC-poor lithofacies vary by up to 2.85‰
and exhibit an overall negative trend from the base of the measured
section at Gwbert Hotel (SN 160 509) from 28.10‰ to 30.50‰ at the
highest OC-poor lithofacies above LH3 (SN 296 530, Fig. 4, Table 2).
Ba/Al values in OC-rich facies vary between 2.5 × 10− 3 and 3.1 × 10− 3
and the mean is higher (mean = 2.8 × 10− 3) than the mean for OC-poor
facies (min. = 1.8 × 10− 3, max. = 4.0 × 10− 3, mean = 2.7 × 10− 3), see Table 4.
Values remain relatively constant through the lower 140 cm of LHl.
There is an abrupt decrease from 3.0 × 10− 3 to 2.7 × 10− 3 at 140 cm
(Fig. 3). Throughout the entire section, Ba/Al ratios and Ba values for
samples from OC-poor lithofacies are statistically lower than in the
OC-rich lithofacies (Table 3) and the correlation between Ba with Al is
good (r2 = 0.6; Fig. 5b). There is no apparent correlation between Ba/Al
and TOC (Fig. 5c).

4.2. Redox proxies

4.2.1. Trace fossils


Sub-sections 1 to 4 (Fig. 6) represent a transect through LH0 at
Aberporth (SN 625 515). Changing burrow diameter reflects changing
ichnogeneric content through the section. Thalassinoides is not
recorded in the LH0 section but is present with Planolites and Chon-
drites elsewhere in the OC-poor lithofacies where BI and maximum
burrow diameter are high (BI = 4, max. burrow diameter = ≥5 mm).
Thalassinoides, Planolites and Chondrites dominate the burrow
mottling of the sediment up to 2.8 m below the base of LH0. At 2.8 m
below the base of LH0 Thalassinoides disappears entirely and Planolites
and Chondrites only are present with BI ≤ 4 and maximum burrow
diameter ≤ 6 mm (sub-sections 1 to 3; Fig. 6). At 4.2 m above the base of
the section (0 cm in sub-section 4, Fig. 6) Chondrites is the only
ichnotaxon present and BI and burrow diameter are correspondingly
low (BI = ≤2, max. burrow diameter ≤ 2 mm). Between 20 and 38 cm in
sub-section 4 and between 22 cm and 38 cm in sub-section 5 strata are
devoid of trace fossils. Also, between the top of sub-section 4 and the
bottom of sub-section 5, no burrowing infauna are recorded. At the top
of sub-section 5 (1.42 m below the top of LH0 Aberporth section 2;
Fig. 6) Planolites is found again corresponding with an increase in
burrow diameters and BI (BI = ≤4, max. burrow diameter = 3 mm).

4.2.2. Geochemical redox proxies; Ni/Co


Fig. 7. Cross-plots for a) Co vs. Ni, b) Ni vs. Al and c) Ni/Al vs. TOC %. Ni values in LHl vary from 10.32 to 92.95 ppm and are typical for
those reported from Phanerozoic shales deposited beneath upwelling
may be regarded as a proxy for arid to humid weathering (Yarincik et al., systems (average shale = 68 ppm Brunsack, 1989; McCann, 1990;
2000). Temple and Cave, 1992; Warnnig and Brumsack, 2000; Boening et al.,
2004; Borchers et al., 2005; Brumsack, 2006, see Table 5), see Table 4.
4. Results Values for samples at 780 cm and 780.75 cm above the base of the LHl
section at Penbryn are anomalously low (10.32 and 14.24 respec-
4.1. Productivity proxies: TOC wt.%,δ13Corg, Ba/Al tively). Ni/Co values fall between 1.03 and 6.69 (mean=2.77), the
mean value for OC-rich LH1 is 2.49 (1.03 to 3.31) whilst values in the
TOC values are higher during deposition of the organic-rich OC-poor facies have a mean of 3.12 (1.37 to 6.69). Crossing the base of
laminated hemipelagite unit LHl. The maximum TOC value of the LHl, Ni/Co values fall abruptly to 1.03 (Fig. 3). Through LHl values
0.61 wt.% in LHl occurs 822.9 cm and 832.5 cm in the measured rise towards the centre (up to 3.31) and then decline towards the
section (mean value for LHl = 0.42 wt.%) whereas the minimum for OC- upper boundary. At 2.5 cm below the top of the LHl, Ni/Co values fall to
poor oxic facies is 0.14 wt.% at 80 cm (mean value for measured 1.4. In the base of the overlying OC-poor lithofacies this value rises
section = 0.31 wt.%, minimum for entire coast section = 0.08 wt.% at Pen abruptly to 6.7.
y Craig, SN 221 252, Fig. 4). Hydrocarbon source rocks typically have Ni/Al values show a different pattern to Ni/Co. Values vary between
N10 wt.% TOC (Tyson, 1995). A weak negative correlation occurs 0.1×10–3 and 6.0×10–4 (mean=2.5 × 10− 4) throughout the same section
between TOC wt.% and δ13Corg (r2 = 0.4), high TOC wt.% values (Fig. 3) whereas in the OC-poor facies values vary from 0.1 × 10− 3 and
correspond to more negative δ13Corg values (Figs. 3 and 5a). 6.0 × 10− 4 (mean=2.3 × 10 − 4). In the OC-rich facies values range from
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 405

Fig. 8. Infaunal biofacies model of (Arthur and Sageman, 1994). Level 1 reflects anaerobic conditions both above and below the SWI, with laminated sediments, the highest levels of
preserved TOC and no evidence of benthic metazoans. At level 2 micro-burrowing disturbs the sedimentary laminae. Level 3 includes small deposit feeding pioneer ichnotaxa such as
Chondrites or Planolites. Level 4 is characterized by an increase in the size and density of Planolites and Chondrites burrows, but no additional ichnotaxa. At level 5 more complex
feeding traces such as Zoophycos and Teichichnus are found. Level 6 is indicated by the appearance of sediment-dwelling ichnotaxa (“domichnia”) such as Thalassinoides. Above level
6 all groups included produce high-diversity assemblages. The Chondrites, Planolites and Thalassinoides ichnocoenoses present in the OC-poor sediments and the transition to OC-rich
sediments represent infaunal biofacies levels 3–4. Interpreted oxygen levels at base of diagram (also from Arthur and Sageman, 1994) correspond to biofacies as shown.

2.1 × 10− 4 and 4.2 × 10− 4 (mean=2.6 × 10− 4). At the base of LH1, Ni/Al deposition. Three distinct biofacies were recognized in the Rhoads
values decrease slightly by 0.21 × 10− 4 (2.67 × 10− 4 to 2.46 × 10− 4). Values and Morse (1971) model; anaerobic, dysaerobic and aerobic, each
gradually rise to a maximum of 5.9 × 10− 4 at the OC-rich OC-poor defined by a specific fauna and characteristic sedimentary fabrics.
boundary at 779.5 cm after which they decrease. In the OC-poor facies, Sageman et al. (1991) defined seven biofacies corresponding to dif-
values rise again fluctuating between 4.3 × 10− 4 (960 cm) and 1.0× 10− 4 ferent levels of oxygenation. Level 1 and 2 indicate anaerobic condi-
(822 cm). Ni shows a good positive correlation with Co (r2 = 0.6; Fig. 7a), tions, level 2–6 dysaerobic conditions and level 7 aerobic conditions
however, Ni is only weakly correlated with Al (r2 = 0.2; Fig. 7b) and TOC (Fig. 8).
wt.% (r2 = 0.2; Fig. 7c). The presence and absence of the three dominant ichnotaxa through
LH0 allows the recognition of three distinct ichnocoenoses that
4.3. Weathering regime proxies, K/Al. Ti/Al correspond to the biofacies levels described by Sageman et al. (1991):
1) Thalassinoides ichnocoenosis (biofacies level 6+), 2) Planolites
During organic-rich sedimentation there is a slight increase in K/Al ichnocoenosis (biofacies level 4–5, and 3) Chondrites ichnocoenosis
values by 0.018 (max. = 0.174, min. = 0.156, OC-rich LH mean = 0.165, OC- (biofacies level 3). Oxygen depletion was initiated prior to the
poor lithofacies mean = 0.163) whereas Ti/Al values do not vary appearance of macroscopic OC, indicated by the disappearance of Tha-
significantly throughout the section (max. = 0.045, min. = 0.40, OC- lassinoides, and declined gradually to fully anoxic conditions 1.4 m into
rich mean = 0.043, OC-poor lithofacies mean = 0.046). Peak values of the event. In the upper part of LH0, O2 levels fluctuated between dysoxic
0.21 and 0.062 for K/Al and Ti/Al respectively are observed just within (Chondrites ichnocoenosis, biofacies level 3) and fully anoxic conditions.
the OC-poor lithofacies at 828.25 cm. Both the means and variances for O2 levels rise through into the overlying OC-poor shales reaching levels
K/Al are not significantly different at a = 0.05 between OC-poor and OC- close to the dysoxic–oxic boundary as marked by the reestablishment of
rich lithofacies. For Ti/Al, only the mean is significantly different at Thalassinoides. Ichnocoenosis boundaries do not coincide with the
α = 0.05 between OC-poor and OC-rich lithofacies see (Table 4). boundaries of the LH suggesting OC burial flux/increased productivity
and redox were not synchronous.
5. Preservation and productivity during deposition of OC-rich During the deposition of LHl we observe, 1) higher TOC wt.% values
laminated hemipelagite which correspond to more negative δ13Corg values when compared to
the OC-poor lithofacies, 2) an increase in the inorganic productivity
5.1. Synchroneity of productivity, preservation and redox proxy Ba/Al and, 3) Ba values in OC-rich facies are higher than those
for OC-poor facies (except samples at 7.84, 9.14 and 10.31 cm), 4) Ba
The trace fossil associations throughout LH0 can be related to values are variable (231 to 758 ppm) though are comparable to values
published models for biofacies and ichnofacies responses in deterior- recorded in modern upwelling systems e.g. Peru margin and Baja,
ating oxygen environments (Rhoads and Morse, 1971; Sageman et al., California (see Table 5) but less than values from euxinic basins e.g.
1991) and can be used to investigate relative timing of anoxia and LH Black Sea (1171 ppm Brumsack, 2006).

Table 5
Trace metal values for modern and Welsh Ordovician OC-rich black shales

Average Peru Namibia Gulf of Nod Glas, Modern Modern Tresaith, Tresaith, Black Sea Black Sea
shale2 margin1 mud lens2 California3 Caradoc, Wales4 turbidites2 abyssal clays2 Wales 16 Wales 26 unit 17 unit 27
Al 16.7 8.9 2.38 8.91 7.69 17.19 14.52 23.4 23.6 6.66 8.86
K 3.8 1.53 0.58 1.66 1.18 2.99 2.28 3.75 3.78 1.26 2.67
Ti 0.78 0.41 0.16 0.39 0.42 0.71 0.7 1.11 1.03 0.28 0.35
Ni 68 74 46 38 208 53.06 47.83 41 74 57 110
Ba 580 314 324 566 228 1427 558.9 558 658 604 1171

Al, K, Ti and Mn in wt.% oxide, all other elements in ppm. lBoening et al. (2004), 2Borchers et al. (2005), 3Brunsack (1989), 4Warnnig and Brumsack (2000), 5Temple and Cave (1992),
6
McCann (1990), 7Brumsack (2006).
406 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

Table 6 Bulk δ13Corg data suggest LHl contains more negative δ13Corg values
End-member values for trace element ratios for depositional redox conditions, Ni/Co than the OC-poor units. We consider this pattern is not attributed to
(Jones and Manning, 1994)
changes in redox because sedimentary data and experimental data
Oxic Dysoxic Suboxic–Anoxic Euxinic studies demonstrate greater depletion of 13Corg in aerobic environments
Ni/Co b 5.00 5.00 7.00 N 7.00 (Lehmann et al., 2002). The difference in carbon isotope depletion
between diagenesis in an aerobic and anaerobic environment is only
0.16‰ (−1.81‰ and −1.65‰ respectively, Lehmann et al., 2002). The
selective loss of different compounds with different reactivities and 13C
These patterns may suggest an increased productivity signal at the values during the decay of organic matter can account for minor shifts
appearance of LHl, however, because Ba values can be obscured by in the 13Corg of marine sediments (Freeman, 2001). However, for algal-
either euxinic conditions (Brumsack, 2006) or diagenesis (Milodowski dominated organic matter, such as in the Ordovician oceans, this effect is
and Zalasiewicz, 1991), it is possible our signal is obscured. Changes in less than 1‰ (Freeman, 2001). Our values are greater than those
productivity at the transition from OC-poor lithofacies to OC-rich LH expected from a change in redox environment and from loss of different
are the most significant because this boundary represents a definite compounds during diagenesis but consistent with increased burial of
12
change in OC-burial and/or productivity. The trace fossil redox data C-enriched carbon. Also, the appearance of OC-rich LH is coincident
from LH0 suggest that euxinic conditions were not achieved (in LH0 at with a decrease in δ13Corg and not with the change in redox inferred in
least) until after 1.8 m across the OC-poor lithofacies OC-rich LHl from the trace fossil distribution of LH0.
boundary. Ba can be considered reliable up until this point in terms
of diagenesis only if similar behaviour in redox is assumed for LHl. A 6. Stability of trace metal proxies
similar pattern in redox to LH0 may be applicable for LHl by inference
from Ba values which decrease approximately 2 m into the OC-rich Both the values and patterns of Ni do not correspond to the sequence
event when oxygenation is inferred to become euxinic and Ba of redox conditions shown by the trace fossils. Indeed, during the
becomes mobile (Fig. 3). Where Ba values subsequently increase deposition of LHl, Ni/Co values are indicative of fully oxic conditions and
dramatically at the OC-rich to OC-poor boundary, may possibly reflect during OC-poor they indicate dysoxia (Table 6). It is therefore clear that
a ‘barite-front’ (Arndt et al., 2006; Brumsack, 2006), a zone of Ni and Co concentrations have been altered during late diagenesis and/
precipitation of barium sulphate at a redox boundary following or low-grade metamorphism and that the trace metals were mobile and
mobilization of Ba under euxinic conditions. can no longer be used as a redox proxy in the studied rocks.

Fig. 9. Light, middle and heavy Rare Earth Elements (La, Sm, Yb) plotted against stratigraphical height above base of LHl, Traeth Penbryn. Sample points are scaled to maximum error.
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 407

On the other hand, we are confident of the reliability of the Ba values The slightly lower K/Al values found in OC-poor facies indicate a
as a productivity proxy because of its enrichment in the OC-rich facies relatively more humid climate with higher fluvial input but not
relative to the Rare Earth elements (REE, La–Yb). The REE have elsewhere significant to develop salinity stratification. As the majority of the Late
been demonstrated to be mobile between turbidite beds and anoxic Katian Welsh Basin deposits are represented by the OC-poor burrow-
hemipelagites (Milodowski and Zalasiewicz, 1991) but REE in the mottled facies, we propose a more humid climate prevailed for much
measured transect of LHl are not enriched or depleted relative to OC- of the late Katian prior to the deposition of the Red Vein.
poor facies (Fig. 9) indicating either no remobilization, remobilization Transgressive shales are formed during relative sea level rise, a
only on a small scale (e.g. cm–dm) or equal loss/gain of REE in each facies. concomitant rise in nutrients/productivity and an upwards expansion
of the OMZ leading to increased preservation potential (Wignall,
7. The origin of Welsh Basin the OC-rich laminate hemipelagites 1991). There are inconsistencies between the Transgressive model and
proxy data for the Welsh Basin LHl. The transgressive black shale
The Runoff model predicts a decrease in products produced under model of Page et al. (2007) is essentially a glacioeustatic variant of the
arid weathering conditions (K relative to Al) and an increase in fluvial Runoff model, i.e. high nutrient input and productivity from glacial
input indicated by higher Ti relative to Al and K. Though not of the meltwater with associated sealevel rise. The stability of our Ti/Al ratio
same magnitude as previous studies utilizing these proxies (e.g. data throughout the OC-poor–LH–OC-poor transitions preclude
Beckmann et al., 2005), we record higher K/Al values and constant Ti/ increased runoff, as previously mentioned. An alternative transgres-
Al values during deposition of LHl suggesting Avalonia experienced a sive black shale model, the ‘Expanding puddle model’ (Wignall, 1991)
relative increase in arid weathering conditions at this time with less involves expansion of the OMZ as sealevel rises. However, in the
fluvial input, making the means to produce a salinity-stratified basin productivity model, sealevel rise alone would only expand the OMZ
from increased continental freshwater runoff unlikely. and raise the pycnocline if accompanied by increased productivity and

Fig. 10. Summary of all proxies during deposition of LH (a) Together, all proxies demonstrate that upwelling and oxygen deterioration commenced prior to the deposition of each LH
event. Falling oxygen conditions were relatively gradual until euxinia was established whilst oxygen recovery was rapid yet fluctuated and occurred prior to the cessation of LH
deposition. (b) Low productivity conditions were typified by weak upwelling and a depressed OMZ whereas high productivity conditions (c) occurred during high wind stress and
increased upwelling that recycled nutrients and 12C-enriched OC. ML = mixed layer.
408 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

oxidation of organic matter which is not implied from sealevel rise.


Implicit in the Transgressive model is increased preservation of OC
(the Preservation model) from decreased thermohaline circulation
and ensuing stagnation of the oceans and/or basin. Such a scenario
could occur during global warming and melting of polar ice which
has not been demonstrated during deposition of the LH. However, the
approximately contemporaneous pre-Hirnantian Boda Event (Fortey
and Cocks, 2005) has been described as a global warming event
from faunal migrations but has in turn been refuted by (Cherns and
Wheeley, 2007) on the grounds of increased global distribution of
cool-water carbonates of Boda Event age.
The relative sea level curve for the Welsh basin (Woodcock, 1990)
is of too low a resolution to determine whether sea level was rising
during the anceps Biozone. An absence of grain-size changes during
the Red Vein suggests sealevel was invariant at the resolution of our
study.
In upwelling systems 12C-enriched CO2 derived from 13C-depleted
recycled organic matter is continually replaced from depth (Pancost
et al., 1999; Peeters et al., 2002), therefore the supply of 12C is
effectively limitless in an open system that has high OC preservation
such as an ocean with permanent deep water anoxia (Leggett, 1980).
Sequestration of this depleted carbon into the marine phyoplankton
through photosynthesis will result in sedimentary organic matter that
is similarly depleted (Pancost et al., 1999). Further to this, in modern
lacustrine (e.g. Hollander and McKenzie, 1991) and marine environ-
ments (e.g. Hofmann et al., 2000; Goericke, 1994; Goericke et al., 1994;
Freeman and Hayes, 1992), δ13Corg has been shown to decrease as the
concentration of dissolved CO2 [CO2aq] in surface waters increases and
as productivity increases. We therefore interpret δ13Corg in the current
setting in terms of a proxy for upwelling intensity as reported from
modern upwelling analogues (e.g. Pancost et al., 1999). Finally, as
recognized by Cave (1965), the presence of phosphate nodules
throughout the succession in both OC-rich LH and OC-poor sediments
could also be a further indication of nutrient-rich upwelling and the
stabilisation of the redox boundary (see also Ingall et al., 1993). Fig. 11. (a) Inferred trade wind positions and directions (from Parrish, 1982) over
Phosphorites have a limited occurrence at the present day, in Avalonia. Palaeogeography adapted from Hermann et al. (2004). (b) Detailed
palaeogeography of the northern margin of Avalonia in the British Isles (after Woodcock
particular being confined to zones of upwelling along west-facing and Strachan, 2000) and inferred position and direction of upwelling.
continental margins (e.g. the Peru margin and S. Africa Ingall et al.,
1993).
The LH unit analysed for inorganic and organic productivity At the present day the strength of the trade winds is related to the
proxies herein (LHl), in light of redox data from LH0, is thus interpreted intensity of the Hadley cell and is related to changes in equatorial high
to represent a period of higher productivity with increased organic pressure. During the Plio-Pleistocene, the position and intensity of the
carbon flux to the seafloor. The amount of marine organic carbon Hadley circulation (and associated trade winds) varied with changes
produced in the photic zone in the modern oceans is dependent in ice volume and solar insolation respectively (Rind, 1998; Ruddiman,
largely on nutrient availability in surface waters (Berger et al., 1989; 2000), the Hadley cell moving polewards during periods of glacial
Jahnke, 1990). We suggest coastal upwelling continually recycled bio- retreat and reduced latitudinal temperature gradient. The intensifica-
limiting nutrients (P, Ba) from depth and 12C-enriched CO2aq. The tion of coastal upwelling in the Welsh Basin and Iapetus Ocean from
elevated Ba levels and negative δ13Corg values observed in LHl are Hadley cell and trade wind position, prior to the Hirnantian glacial
consistent with being the product of increased productivity from maximum, may therefore reflect latitudinal movement of the Hadley
upwelling (Fig. 10a, c). During these periods of high productivity, cell during an increase (northward movement of the Hadley cell) or
increased oxidative respiration of organic matter resulted in; 1) the decrease (southward movement of the Hadley cell) in the size of the
seafloor becoming increasingly oxygen depleted and, 2) an upwards Upper Ordovician ice sheet, depending on where the Hadley cell was
expansion of the OMZ. Organic-poor shales were deposited during initially positioned relative to Avalonia. Subsequent short-term cyclic
periods of decreased productivity and organic carbon flux and a return patterns of increased upwelling (possibly represented by LH0–LH3)
to more oxygenated seafloor ensued (Fig. 10b). could reflect orbitally-moderated changes in solar insolation and ice
Coastal upwelling intensity can vary with a number of factors, volume.
but the most important of these is the strength of the offshore
winds (Hay and Brock, 1992). During the late Katian, global 8. Conclusions
atmospheric circulation models (Parrish, 1982) and palaeoconti-
nental configurations (Hermann et al., 2004) infer that the Welsh The fine grained siliciclastic rocks of the Welsh Basin record an
Basin could have been located beneath the zone of southwesterly evolving sequence of events during the transition from Ordovician
prevailing trade winds on a northwestwards facing coast (Fig. 11). greenhouse to icehouse climate states. We postulate that the proxy
We hypothesize that the cyclic nature of black shale deposition in records from one OC-rich laminated hemipelagite unit (LHl) presented
the Red Vein reflects the changing intensity of the upwelling system herein indicate short-lived, cyclic enhanced coastal upwelling and
in response to the changing strength of the offshore trade winds surface organic productivity along the southern Iapetus margin.
(see also Cave, 1965). Increased upwelling and productivity could have been induced by
T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410 409

increased wind stress from the intensification of coast-parallel to Cave, R., 1965. The Nod Glas sediments of Caradoc age in North Wales. Geological
Journal 4, 279–298.
offshore SE trade winds. Higher productivity and OC supply to the Cherns, L., Wheeley, J.R., 2007. A pre-Hirnantian Late Ordovician interval of global
deep ocean would likely have resulted in a rise in the position of the cooling the Boda Event re-assessed. Palaeogeography, Palaeoclimatology, Palaeoe-
OMZ, which breached the semi-restricted Welsh Basin, in response to cology 251 (3–4), 449–460.
Cocks, L.R.M., McKerrow, W.S., Van Staal, C.R., 1997. The margins of Avalonia. Geological
increased oxygen depletion of the water column. We hypothesize that Magazine 134, 627–636.
black shale deposition may have been controlled by complex climate– Davies, J.R., Fletcher, C.J.N., Waters, R.A., Wilson, D., Woodhall, D.G., Zalaziewicz, J.A.,
ocean–biosphere interactions possibly reflecting responses to ice 1997. Geology of the country around Llanilar and Rhayader. Memoir of the British
Geological Survey, Sheets 178 and 179 (England and Wales). The Stationary Office,
volume changes during the developing end-Ordovician glaciation.
London, p. 267.
Replication of these results from other OC-rich units of similar age in Davies, J.R., Waters, R.A., Wilby, P.R., Williams, Wilson, D., 2003. Geology of the Cardigan
the Welsh Basin (e.g. LH0, LH2 and LH3) is necessary to further support and Dinas Island district a brief explanation of the geological map. Sheet
explanation of the British Geological Survey. 1 : 50000 Sheet 193 (including part
this interpretation. Metamorphism to anchizone metamorphic facies
of Sheet 210) Cardigan and Dinas Island (England and Wales). British Geololical
reduces the number of ocean–environment proxies that can be a used Survey, p. 26.
to elucidate palaeoenvironmental conditions. We consider the Davies, J.R., Sheppard, T.H., Waters, R.A., Wilson, D., 2006. Geology of the Llangranog
productivity proxies, Ba/Al, TOC %, δ13Corg and the climate-sensitive district a brief explanation of the geological map. Sheet explanation of the British
Geological Survey. 1 : 50000 Sheet 194 (England and Wales). British Geololical
detrital input proxies K/Al and Ti/Al to be robust and could be widely Survey, p. sw.
applied in the slate grade rocks commonly found in orogenic belts. De La Rocha, C.L., 2004. The biological pump. In: Elderfield, H. (Ed.), The Oceans
and Marine Geochemistry. Vol. 6 of Treatise on Geochemistry. Elsevier Pergamon,
pp. 83–112.
Acknowledgments Dypvik, H., 1984. Geochemical composition and depositional conditions of the Upper
Jurassic and Lower Cretaceous Yorkshire clays, England. Geological Magazine 121,
T. J. Challands would like to acknowledge the receipt of a University 489–504.
Ekdale, A.A., Mason, T., 1988. Characteristic trace-fossil associations in oxygen-poor
of Durham Postgraduate Scholarship and a British Geological Survey sedimentary environments. Geology 16, 720–723.
British Universities Funding Initiative grant. Comments by P. B. Fettes, D.J., Long, C.B., Bevins, R.E., Max, M.D., Olover, G.J.H., Primmer, T.J., Thomas, L.J.,
Wignall and J. A. Zalasiewicz have helped to greatly improve the Yardley, B.W.D., 1985. Grade and time of metamorphism in the Caledonide Orogen
of Britain and Ireland. In: Harris, A.L. (Ed.), The nature and timing of orogenic
manuscript. J. R. Davies and D. Wilson publish with the permission of activity in the Caledonian rocks of the British Isles. Blackwell for the Geological
the Executive Director, British Geological Survey (NERC). Society of London, U.K.
Fortey, R.A., Cocks, L.R.M., 2005. Late Ordovician global warming — the Boda event.
Geology 33 (5), 405–408.
References
Fortey, R.A., Harper, D.A.T., Ingham, J.K., Owen, A.W., Parkes, M.A., Rushton, A.W.A.,
Woodcock, N.H., 2000. A Revised Correlation of Ordovician rocks in the British Isles.
Armstrong, H.A., Coe, A.L., 1997. Deep-sea sediments record the geophysiology of the Geological Society of London Special Report, vol. 24. 83 pp.
late Ordovician glaciation. Journal of the Geological Society, London 154, 929–934. Freeman, K.H., 2001. Reviews in mineralogy and geochemistry. Isotopic biogeochem-
Armstrong, H.A., Owen, A.W., 2002. Euconodont palaeobiogeography and the closure of istry of marnie organic carbon, vol. 43. Mineralogical Society of America and
the Iapetus Ocean. Geology 30 (12), 1091–1094. Geochemical Society, Washington, DC, United States, pp. 579–605. Ch. 11.
Arndt, S., Brumsack, H.-J., Wirtz, K.W., 2006. Cretaceous black shales as active Freeman, K.H., Hayes, J.M., 1992. Fractionation of carbon isotopes by phytoplankton and
bioreactors; a biogeochemical model for the deep biosphere encountered during estimates of ancient CO2 levels. Global Biogeochemical Cycles 6, 185–198.
ODP Leg 207 (Demerara Rise). Geochimica et Cosmochimia Acta 70 (2), 408–425. Goericke, R., 1994. Variations in marine δ13C with latitude, temperature and dissolved
Arthur, M., Sageman, B., 1994. Marine black shales: depositional mechanisms and CO2 in the world ocean. Global Biogeochemical Cycle 8, 85–90.
environments of ancient deposits. Annual Review of Earth and Planetary Sciences Goericke, R., Montoya, J.P., Fry, B., 1994. Physiology of isotope fractionation in algae and
22, 499–551. cyanobacteria. In: Lajtha, K., Michener, B. (Eds.), Stable isotopes in ecology.
Beckmann, B., Flögel, S., Hofmann, P., Schulz, M., Wagner, T., 2005. Orbital forcing of Blackwell Scientific, Boston, pp. 199–233.
Cretaceous river discharge in tropical Africa and ocean response. Nature 437 (8), Hannigan, R.E., Basu, A., 1998. Late diagenetic trace element remobilization in organic-
241–244. rich black shales of the Taconic foreland basin of Quebec, Ontario and New York. In:
Berger, W.H., Smetacek, V.S., Wefer, G., 1989. Productivity of the Oceans: Present and J. Schieber, W.Z., Sethi, P.S. (Eds.), Shales and Mudstones, vol. II, pp. 209–233.
Past. Wiley Interscience, New York, p. 471. Hay, W.W., Brock, J.C., 1992. Temporal variation in intensity of upwelling off southwest
Berner, R.A., 1994. Geocarb II: a revised model for atmospheric CO2 over Phanerozoic Africa. In: Summerhayes, C.P., P.W.L., K.C., Emeis (Eds.), Upwelling Systems:
time. American Journal of Science 294, 56–91. evolution since the Early Miocene. Geological Society Special Publication, vol. No.
Berner, R.A., Kothavala, Z., 2001. Geocarb III: a revised model for atmospheric CO2 over 64, pp. 463–497.
Phanerozoic time. American Journal of Science 301 (2). Hermann, A.D., Patzkowsky, M.E., Pollard, D., 2004. The impact of paleogeography,
Berry, W.B.N., Wilde, P., 1978. Progressive ventilation of the oceans; an explanation for pCO2, poleward ocean heat transport and sea-level change on global cooling during
the distribution of the lower Paleozoic black shales. American Journal of Science the Late Ordovician. Palaeogeography, Palaeoclimatology, Palaeoecology 206,
278 (3), 257–275. 59–74.
Bevins, R.E., Merriman, R.J., 1988. Compositional controls on coexisting prehnite Herrmann, A.D., Haupt, B.J., Patzkowsky, M.E., Seidov, D., Slingerland, R.L., 2005.
actinolite and prehnite pumpellyite facies assemblages in the TalY-Fan metabasite Response of Late Ordovician paleoceanography to changes in sea level, continental
intrusion, North Wales: implications for Caledonian metamorphic field gradients. drift and atmospheric pCO2: potential causes for long-term cooling and glaciation.
Journal of Metamorphic Geology 6, 17–39. Palaeogeography, Palaeoclimatology, Palaeoecology 210, 385–401.
Bevins, R.E., Robinson, D., 1988. Low grade metamorphism in the Welsh Basin Lower Hofmann, M., Wolf-Gladrow, D.A., Takashashi, T., Sutherland, S.C., Six, K.D., Maier-
Palaeozoic succession: An example of diastathermal metamorphism? Journal of the Reimer, E., 2000. Stable carbon isotope distribution of particulate organic matter in
Geological Society, London 145, 363–366. the ocean: a model study. Marine Chemistry 72, 131–150.
Bjerrum, C.J., Bendtsen, J., Legarth, J.J.F., 2006. Modeling organic carbon burial during Hollander, D.J., McKenzie, J.A., 1991. CO2 control on carbon isotope fractionation during
sea level rise with reference to the Cretaceous. Geochemistry, Geophysics, aqueous photosynthethesis: a palaeo-pCO2 barometer. Geology 19, 929–932.
Geosystems G3 7 (5). doi:10.1029/2005GC001032. Ingall, E.D., Bustin, R.M., Van Cappellen, P., 1993. Influence of water column anoxia on
Boening, P., Brumsack, H.-J., Boettcher, M.E., Schnetger, B., Kriete, C., Borchers, S.L., 2004. the burial and preservation of carbon and phosphorus in marine shales. Geochimica
Geochemistry of Peruvian near-surface sediments. Geochimica et Cosmochimia et Cosmochimia Acta 57 (2), 303–316.
Acta 68 (21), 4429–4451. Jahnke, R.A., 1990. Ocean flux studies: a status report. Review of Geophysics 28,
Borchers, S.L., Schnetger, B., Böning, P., Brunsack, H.-J., 2005. Geochemical signatures of 381–398.
the Namibian diatom belt: perennial upwelling and intermittent anoxia. Geochem- Jones, B., Manning, D.A.C., 1994. Comparison of geochemical indices used for the
istry, Geophysics, Geosystems G (super 3) 6 (6). doi:10.1029/2004GC000886. interpretation of palaeoredox conditions in ancient mudstones. Chemical Geology
Bottrell, S.H., Greenwood, P.B., Yardley, B.W.D., Shepherd, T.J., Spiro, B., 1990. 111 (1–4), 111–129.
Metamorphic and post-metamorphic fluid-flow in the low-grade rocks of the Kuypers, M.M.M., Pancost, R.D., Nijenhuis, I.A., Sinninghe Damsté, J.S., 2002.
Harlech Dome, North Wales. Journal of Metamorphic Geology 8, 131–143. Enhanced productivity led to increased organic carbon burial in the euxinic
Brunsack, H.-J., 1989. Geochemistry of Recent TOC-rich sediment from the Gulf of North Atlantic basin during the late Cenomanian oceanic anoxic event.
California and the Black sea. Geologische Rundschau 78 (3), 851–882. Paleoceanography 17 (4).
Brumsack, H.-J., 2006. The trace metal content of recent organic carbon-rich sediments: Kuypers, M.M.M., Lourens, L.J., Rijpstra, W., Irene, C., Pancost, R.D., Nijenhuis, I.A.,
implications for Cretaceous black shale formation. Palaeogeography, Palaeoclima- Damste, J.S., 2004. Orbital forcing of organic carbon burial in the proto-North
tology, Palaeoecology 232, 344–361. Atlantic during oceanic anoxic event 2. Earth and Planetary Science Letters 228 (3–4),
Canfield, D.E., 1989. Sulfate reduction and oxic respiration in marine sediments: 465–482.
implications for organic carbon preservation in euxinic environments. Deep-sea Leggett, J.K., 1980. British Lower Palaeozoic black shales and their palaeo-oceanographic
Research 36, 121–138. significance. Journal of the Geological Society, London 137, 139–156.
410 T.J. Challands et al. / Palaeogeography, Palaeoclimatology, Palaeoecology 273 (2009) 395–410

Lehmann, M.F., Bernasconi, S.M., Barbieri, A., McKenzie, J.A., 2002. Preservation of Roberts, B., Merriman, R.J., Hirons, S.R., Fletcher, C.J.N., Wilson, D., 1996. Synchronous
organic matter and alteration of its carbon and nitrogen isotope composition during very low-grade metamorphism, contractions and inversion in the central part of
simulated and in situ early sedimentary diagenesis. Geochimica et Cosmochimia the Welsh lower Palaeozoic basin. Journal of the Geological Society, London 153,
Acta 66 (20), 3573–3584. 277–285.
Lev, S.M., Filer, J.K., Tomascak, P., 2008. Orogenesis vs. diagenesis: can we use organic- Robinson, D., Bevins, R.E., 1986. Incipient metamorphism in the Lower Palaeozoic
rich shales to interpret the tectonic evolution of a depositional basin? Earth Science marginal basin of Wales. Journal of Metamorphic Petrology 4, 101–113.
Reviews 86 (1–4), 1–14. Ruddiman, W.F., 2000. Earth's climate: past and future. W.H. Freeman and Company,
Löwemark, L., Schönfeld, J., Werner, F., Scäfer, P., 2004. Trace fossils as a paleoceano- New York.
graphic tool: evidence from late Quaternary sediments of the southwestern Iberian Sageman, B.B., Wignall, P.B., Kauffman, E.G., 1991. Biofacies models for organic-rich
margin. Marine Geology 204, 27–41. facies: tool for paleoenvironmental analysis. In: G. Einsele, W.R., Seilacher, A. (Eds.),
Maslin, M.A., Swann, G.E.A., 2005. Isotopes in Palaeoenvironmental Research. Ch.6 Cycles and Events in Stratigraphy. Springer-Verlag, Berlin, pp. 542–564.
Isotopes in marine sediments. Springer, pp. 227–269. Savrda, C.E., 1995. Ichnologic applications in paleoceanographic, paleoclimatic and sea-
McCann, T., 1990. Distribution of Ordovician–Silurian ichnofossil assemblages in Wales level studies. Palaios 10, 565–577.
implications for Phanerozoic ichnofaunas. Lethaia 23, 243–255. Savrda, C.E., Bottjer, D.J., 1994. Ichnofossils and ichnofabrics in rythmically bedded
Melchin, M.J., Holmden, C., Williams, S.H., 2003. Correlation of graptolite zones, pelagic/hemipelagic carbonates: recognition and evaluation of benthic redox and
chitinozoan biozones, and carbon isotope curves through the Hirnantian. In: scour cycles. In: Orbital forcing and Cyclic Sequences. International Association of
Albanesi, G.L., B.M.S., S.H., Peralta (Eds.), Ordovician from the Andes Proceedings of Sedimentologists, Special Publication, pp. 195–210.
the 9th International Symposium on the Ordovician System. INSUGEO, Serie Schofield, D.I., Davies, J.R., Waters, R.A., Wilby, P.R., Williams, M., Wilson, D., 2004.
Correlación Geológica, vol. 17, pp. 101–104. Geology of the Builth Wells district a brief explanation of the geological map. Sheet
Meyers, P.A., Arnaboldi, M., 2005. Trans-Mediterranean comparison of geochemical Explanation of the British Geological Survey. 1:50 000 sheet 196 Builth Wells
productivity proxies in a mid-Pleistocene interrupted sapropel. Palaeogeography, (England and Wales).
Palaeoclimatology, Palaeoecology 222, 313–328. Shimmield, G.B., 1992. Can sediment geochemistry record changes in coastal upwelling
Milodowski, A.E., Zalasiewicz, J.A., 1991. Redistribution of rare earth elements during palaeoproductivity? In: Summerhayes, C.P., Prell, W.L., Emeis, K.C. (Eds.), Upwelling
diagenesis of turbidite/hemipelagite mudrock sequences of Llandovery age from systems: Evolution since the Eary Miocene. . Geological Society Special Publication,
central Wales. In: Morton, A.C., Todd, S.P., Houghton, P.D. (Eds.), Developments in vol. 64. The Geological Society, London, pp. 29–46.
sedimentary provenance studies. Geological Society Special Publication, vol. 57. Sobarzo, J., Bravo, L., Donoso, D., Garcés-Vargas, J., Schneider, W., 2007. Coastal
Geological Society, London, pp. 101–124. upwelling and seasonal cycles that influence the water column over the continental
Nara, F., Tani, Y., Soma, Y., Soma, M., Naraoka, H., Watanabe, T., Horiuchi, K., Kawai, T., shelf off central Chile. Progress in Oceanography 75 (3), 363–382.
Oda, T., T., N., 2005. Response of phytoplankton productivity to climate change Stow, D.A.V., Wetzl, A., 1987. Hemiturbidite: a new type of deep-water sediment.
recorded by sedimentary photosynthetic pigments in Lake Hovsgol (Mongolia) for Proceedings of the Ocean Drilling Program, distal Bengal Fan; covering Leg 116 of
the last 23,000 years. Quaternary International 136, 71–81. the drilling vessel JOIDES Resolution, Colombo, Sri Lanka, to Colombo, Sri Lanka,
Ottley, C.J., Pearson, D.G., Irvine, G.J., 2003. A routine method for the dissolution of sites 717–719, 2 July 1987–19 August 1987. Texas A and M University, Ocean Drilling
geological samples for the analysis of REE and trace elements via ICP-MS. In: Program, United States, vol. 116, pp. 25–34.
Holland, J.G., Tanner, S.D. (Eds.), Plasma Source Mass Spectrometry: Applications Temple, J.T., Cave, R., 1992. Preliminary report on the geochemistry and mineralogy
and Emerging Technologies. Royal Society of Chemistry, Cambridge, pp. 221–230. of the Nod Glas and related sediments (Ordovician) of Wales. Geological Magazine
Page, A., Zalaziwicz, J., Williams, M., Popov, L., 2007. Were transgressive black shales a 129 (5), 589–594.
negative feedback modulating glacioeustacy in the Early Palaeozoic icehouse? In: Tribovillard, N., Ramdani, A., Trentesaux, A., 2005. Controls on organic accumulation in
Williams, M., H.A.M.G.F.J., D.N., Schmidt (Eds.), Deep-Time Perspectives on Climate upper Jurassic shales of northwestern Europe as inferred from trace-metal geochem-
Change: Marrying the Signal from Computer Models and Biological Proxies. The istry. In: Harris, N.B. (Ed.), The deposition of organic-carbon-rich sediments: models,
Micropalaeontological Society Special Publications. The Geological Society, London, mechnisms and consequences. Society for Sedimentary Geology Special Publication,
pp. 123–156. vol. 82, pp. 145–164.
Pancost, R.D., Freeman, K.H., Wakeham, S.G., 1999. Controls on the carbon-isotope Twichell, S.C., M.P.A., Diester-Haass, L., 2002. Significance of high C/N ratios in organic
compositions of compounds in Peru surface waters. Organic Geochemistry 30 (5), carbon-rich Neogene sediments under the Benguela Current upwelling system.
319–340. Organic Geochemistry 33, 715–722.
Parrish, J.T., 1982. Upwelling and petroleum source beds, with reference to Palaeozoic. Tyson, R.V., 1995. Sedimentary organic Matter: Organic Facies and Palynofacies.
AAPG Bulletin 66, 750–774. Chapman and Hall, London.
Peacor, D.R., Voveney, R.M.J., Zhao, G., 2000. Authigenic illite and organic matter; the Tyson, R.V., 2001. Sedimentation rate, dilution, preservation and total organic carbon:
principal hosts of vanadium in the Mecca Quarry Shale at Velpen, Indiana. Clay and some results of a modelling study. Organic Geochemistry 32 (2), 333–339.
Clay Minerals 48 (3), 311–316. Underwood, C.J., Crowley, S.F., Marshall, J.D., Brenchley, P.J., 1997. High-resolution
Peeters, F.J.C., Brummer, G.-J.A., Ganssen, G., 2002. The effect of upwelling on the carbon isotope stratigraphy of the basal Silurian Stratotype (Dob's Linn, Scotland)
distribution and stable isotope composition of Globigerina bulloides and Globigeri- and its global correlation. Journal of the Geological Society, London 154, 709–718.
noides ruber (planktic Foraminifera) in modern surface waters of the NW Arabian Vandenbroucke, T.R.A., Vanmeirhaeghe, J., 2007. Acta palaeontologica Sinica 46,
Sea. In: Ganssen, G. (Ed.), From process studies to reconstruction of the 497–501.
palaeoenvironment; advances in palaeoceanography, Global and Planetary Change, Vilinski, J.C., Domack, E., 1998. Temporal changes in sedimentary organic carbon from
vol. 34. Elsevier, Amsterdam, Netherlands, pp. 269–291. the Ross Sea Antarctica: inferred changes in ecosystems and climate. Eos
Peters-Kottig, W., Strauss, H., Kerp, H., 2006. The land plant delta 13C record and plant (Transactions, American Geophysical Union) 79, 157.
evolution in the late Palaeozoic. Palaeogeography, Palaeoclimatology, Palaeoecol- Warnnig, B., Brumsack, H.-J., 2000. Trace metal signatures of eastern Mediterranean
ogy 240 (1–2), 237–252. sapropels. Palaeogeography, Palaeoclimatology, Palaeoecology 158 (3–4), 293–309.
Pope, M.C., Steffen, J.B., 2003. Widespread, prolonged late Middle to Late Ordovician Webby, B.D., Droser, M.L., Paris, F., Percival, I., 2004. The Great Ordovician Biodiversi-
upwelling in North America: a proxy record of glaciation? Geology 33 (6), 214–215. fication Event. Columbia University Press, New, p. 484.
Prakash Babu, C., Brumsack, H.-J., Schnetger, B., Boettcher, M.E., 2002. Barium as a Wignall, P., 1991. Model for transgressive black shales? Geology 19, 167–170.
productivity proxy in continental margin sediments; a study from eastern Arabian Wignall, P.B., 1994. Black shales. Clarendon Press, Oxford.
Sea. Marine Geology 184 (3–4), 189–206. Woodcock, N.H., 1990. Sequence stratigraphy of the Palaeozoic Welsh Basin. Journal of
Pugh, W.J., 1923. The geology of the district around Corris and Aberllefenni the Geological Society, London 147, 537–547.
(Merionethshire). Quarterly Journal of the Geological Society, London 85, 242–306. Woodcock, N.H., Strachan, R., 2000. Geological history of Britain and Ireland. Blackwell,
Rhoads, D.C., Morse, J.W., 1971. Evolutionary and ecologic significance of oxygen p. 423.
deficient marine basin. Lethaia 4, 413–428. Yapp, C.J., Poths, H., 1992. Ancient atmospheric CO2 pressures inferred from natural
Rind, D., 1998. Latitudinal temperature gradients and climate change. Journal of goethites. Nature 353, 342–344.
Geophysical Research 103 (D6), 5943–5971. Yarincik, K.M., Murray, R.W., Peterson, L.C., 2000. Climatically sensitive eolian and
Roberts, B., Merriman, R.J., Pratt, W., 1991. The influence of strain, lithology and hemipelagic deposition in the Cariaco Basin, Venezuela, over the past 578,000 years.
stratigraphical depth on white mice (Illite) crystallinity in mudrocks from the Results from Al/Ti and K/Al. 15, 210–228.
vicinity of the Corris Slate Belt, Wales: implications for the timing of metamorph-
ism in the Welsh Basin. Geological Magazine 128, 633–645.

You might also like